A CESIUM - 133 NMR STUDY OF THE STAT'ICS AND ‘ DYNAMICS OF CESIUM ION COMPLEXATION BY CROWNS AND CRYPTANDS IN VARIOUS SOLVENT S Dissertation for the Degree of Ph. D. MICHIGAN STATE UNIVERSITY ELIZABETH HUN-I MEI 1977 3 1293 01073 2067 ,b'n'fimuu - " LIBRAR L Y N" 1 . A C "cm. I YIJIObun 3"?” “L (“WWW ——— This is to certify that the thesis entitled A CESIUM-133 NMR STUDY OF THE STATICS AND DYNAMICS OF CESIUM ION COMPLEXATION BY CROWNS AND CRYPTANDS IN VARIOUS SOLVENTS. presented by ELIZABETH HUN-I MEI has been accepted towards fulfillment - of the requirements for _Eh._D..__ degree in Lhamisiry c» y/f’fl “/4 wbjéz/x/ Major profes Date NOV. LI', 1976 0-7 639 J, g?! 14 “3 .081“ 1 '9 23 {ECI I " aécoq A v-“fil A CESIUM-l CESIUM ICI ABSTRACT A CESIUM-l33 NMR STUDY OF THE STATICS AND DYNAMICS OF CESIUM ION COMPLEXATION BY CROWNS AND CRYPTANDS IN VARIOUS SOLVENTS BY Elizabeth Hun-I Mei Chemical shifts of the cesium-133 nucleus were measured in six nonaqueous solvents relative to 0.5 M aqueous cesium bromide. Cesium tetraphenylborate (CsTPBT triiodide and thiocynate were used to determine the infinite dilution chemical shifts in pyridine (PY), propylene carbonate (PC), dimethylformamide (DMF), dimethylsulfoxide (DMSO), aceto- nitrile (MeCN), and acetone. The corresponding ion-pair formation constants were determined from chemical shift concentration data with the aid of a weighted nonlinear least squares program (KINFIT). The association constant for CsSCN in pyridine is 9001200 while for CsTPB in pyridine it is 370120, in PC it is 1617, in MeCN it is 40:10, and in acetone it is 2213. The uncertainties given are standard deviation estimates. Cesium-133 NMR studies were also performed on cesium tetraphenylborate complexes with five ligands in the six nonaqueous solvents mentioned. These ligands were 18-Crown-6 I18C6), dibenzo- tad-€222 (C222) have different t cotplexation abi 1:1 and 2:1 (lig solvents with th 103. A new £le to analyze data I2le formation. fected by the ge It IIas also show in the equilibri Complexation by Elizabeth Hun-I Mei (18C6), dibenzo-18C6 (DBC), dicyclohexyl-18C6 (DCC), cryp- tand-C222 (C222), and monobenzo-C222 (C2228). These ligands have different topologies and substituents which affect the complexation ability. Cesium tetraphenylborate forms both 1:1 and 2:1 (ligand/Cs+) complexes with 18C6 in all six solvents with the first formation constant (K1) larger than 103. A new EQN subroutine of the KINFIT program was written to analyze data which show both 1:1 and sandwich complex (2:1) formation. It was found that both K1 and K2 are af- fected by the geometry and substituents of the crown ligands. It was also shown that the solvent plays an important role in the equilibrium process. For example, the K values for 8, K = 71:1, 2 in PC K1 = (1.5t0.6) x 104, K2 = 8:2, in acetone K1 > 107. x2 = 3410.5, in our K1 = (9:3) x 103, K2 = 2.44:0.05, in DMSO K complexation by 18C6 in pyridine are 2K1>10 1 = (1.1:0.1) x 103, K2 = (110.4) and in MeCN K1 > 105, K2 = 4.4:0.3. The attachment of a substituent on the ring of 18C6 yielded values of K1 in the order 18C6 > DCC > DBC. However, probably because of steric effects, the K2 values are in the order DBC > 18C6 > DCC (at least in pyri- dine). A thorough study of 18C6 complexes with CsTPB in pyri— dine was made at various temperatures (from 25° to -44°C). For the purpose of this study a new temperature independent reference was designed. Its validity was tested and it was used to show that ion-ion and ion-solvent interactions give temperature-dependent chemical shifts. The values of the first format too large to be c were determined changes for the 152 = -6.2:0.1 K 152 = -ll.2:0.3 reaction of (35+. KCdl/mole. The fomati Vere also obtain t . *Cmatlon consta solvents . For i x102 (Dm) ' (27 2228, with a bet: weaker complexes ZPYI to Zero (DH! a? 45 ‘endent studie F . o. kinetics B ' C i n the Solution A; ”- th e formation 0f CS+ Elizabeth Hun-I Mei the first formation constant at various temperatures were too large to be determined by NMR techniques but K2 values were determined and used to obtain the enthalpy and entropy changes for the second complexation step. The results are: AH2 = -6.210.l Kcal/mole (A63) = -2.8310.004 Kcal/mole, 298 A82 = -ll.210.3 e.u. A kinetics study of the decomplexation reaction of Cs+-18C6 gave an activation energy of 810.3 Kcal/mole. The formation constants of C222 and C2228 complexes were also obtained from the NMR chemical shift data. The formation constants showed the same trends with various solvents. For instance, K1 values for C222 are >105 (PY), 3 (pc), (10.810.8) x 103 (acetone), (1.510.1) 4 (1011) x 10 x 102 (one), (2713) (DMSO), and (411) x 10 (MeCN). Cryptand— 2228, with a benzo group on one of the ether chains forms weaker complexes, with K values ranging from (5.710.8) x 103 (FY) to zero (DMSO). Chemical shift-mole ratio temperature dependent studies were also carried out, as well as studies of kinetics. Both gave evidence for two types of complexes in the solution. The results are interpreted on the basis of the formation of both inclusive and exclusive complexes of Cs+ by C222. Enthalpies and entropies of formation were also calculated by using the KINFIT program and it was found that both quantities are sensitive to the solvent for the complexation of free cesium ions to form the exclusive complex. The conversion of the exclusive to the inclusive complex is much less sensitive to solvent. The activation energy (Ba) for t I CsTPB in PC is 1~ obtained from kit are Ea = 8.510.5 (for 18C6) . It 5 restricted geome° the largest acti~ ion. Elizabeth Hun-I Mei energy (Ea) for the complexation reaction of C2228 with CsTPB in PC is 14:0.6 Kcal/mole. By comparison the values obtained from kinetics studies of crown complexes in PC are Ea = 8.5i0.5 Kcal/mole (for DCC) and E3 = 814 Real/mole (for 18C6). It appears that the higher rigidity and restricted geometry of the ligand combine to give C2228 the largest activation energy for removal of a cesium ion. A CESIUM ornnnn CROWN S 1n Par+ A CESIUM-l33 NMR STUDY OF THE STATICS AND DYNAMICS OF CESIUM ION COMPLEXATION BY CROWNS AND CRYPTANDS IN VARIOUS SOLVENTS BY ELIZABETH HUN-I MEI A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirement for the degree of DOCTOR OF PHILOSOPHY Department of Chemistry 1977 To My Parents. ii The author Professors Jan guidance, enc< out this study Gratitude Michigan State Ildmirtistratior, financial aid. Special th brother Dr. Ed meflt and SuggeI to the memberd and PrOfeSSOr moral Support, Natty thaw Frank Bennie. A eter in operaq I am deepl love and Cordi I dedicate thi ACKNOWLEDGMENTS The author wishes to express her sincere gratitude to Professors James L. Dye and Alexander I. Popov for their guidance, encouragement and whole-hearted support through- out this study. Gratitude is also extended to the Department of Chemistry, Michigan State University, the 0.8. Energy and Development Administration and the National Science Foundation for financial aid. Special thanks go to my Uncle Dr. Tony S. Shen and my brother Dr. Howell H. Mei for their constant encourage- ment and suggestions in many ways. Thanks are also extended to the members of the research groups of both Professor Dye and Professor Popov for their constant stimulation and moral support. Many thanks also go to Mr. Wayne Burkhardt and Mr. Frank Bennis for their efforts in keeping the NMR spectrom— eter in operating condition. I am deeply grateful to my parents for their great love and cordial concern throughout all my life. To them I dedicate this thesis. iii Chapter LIST OF TABLES LIST 01’ FIGURE omens I. n: l. INTROU II. HISTOE 3- Ma P 3 Co Ma C Nu (j (1‘ (ii III CONCLr AND L] II' SOLVEe PREPAI III, THE: N» AND D}: EESIUM is VOMPLBXES ARIOU So INTRO); I1_ . bIINVEST Chapter LIST OF LIST OF CHAPTER 1. II. III. CHAPTER I. II. III. CHAPTER TABLE OF CONTENTS TABLES. . . . . . . . . . . . . . . . FIGURES . . . . . . . . . . . . . . . I. HISTORICAL. . . . . . . . . . . . INTRODUCTION. . . . . . . . . . . . . HISTORICAL BACKGROUND . . . . . . . . A. Macrocyclic Polyether-Crown Com- plexation of Alkali Metal Ions. . 8. Complexation of Metal Ions by Macroheterocyclic Ligands-Cryptands C. Nuclear Magnetic Resonance. . . . (i) Introduction . . . . . . . . (ii) Chemical Shift Studies of Electrolyte Solutions. . . . (iii) Cesium Nuclear Magnetic Resonance. . . . . . . . . . CONCLUS IONS . O C O O C C O O O O O C II. EXPERIMENTAL PART. . . . . . . . SYNTHESIS OF CESIUM TETRAPHENYLBORATE AND LIGAND PURIFICATION . . . . . . . SOLVENT PURIFICATION AND SAMPLE PREPARATION . . . . . . . . . . . . . THE NMR SPECTROMETER; MEASUREMENT AND DATA HANDLING . . . . . . . . . . III. STUDY OF FORMATION CONSTANTS OF CESIUM TETRAPHENYLBORATE ION PAIR AND OF COMPLEXES WITH MACROCYCLIC LIGANDS IN VARIOUS SOLVENTS. . . . . . . . . . . . . I. II. INTRODUCTION. . . . . . . . . . . . . INVESTIGATION OF CESIUM SALTS IN NONAQUEOUS SOLVENT s o o o o o o o o 0 iv Page vii xiv 15 25 25 27 38 43 44 45 47 52 57 58 59 Chapter III. FORMA' PHENYI AND Cl IV. TEMPE? SHIFT omen. Chapter III. FORMATION CONSTANTS OF CESIUM TETRA- PHENYLBORATE COMPLEXES WITH cnowns AND CRYPTANDS IN VARIOUS SOLVENTS . . . . (i) Complexation Reactions with Crowns . . . . . . . . . . . . . (ii) Complexation Reactions with Cryptands. . . . . . . . . . . . IV. TEMPERATURE DEPENDENCE OF THE CHEMICAL SHIFT AND ITS RELATIONSHIP TO THE THERMO- DYNAMICS OF COMPLEXATION REACTIONS. . . . A. CESIUM TETRAPHENYLBORATE COMPLEXES WITH 18C6 IN PYRIDINE . . . . . . . . B. CESIUM TETRAPHENYLBORATE COMPLEXES WITH C222 IN VARIOUS SOLVENTS . . . . CHAPTER IV. A STUDY OF THE DYNAMICS OF CESIUM TETRAPHENYLBORATE COMPLEXES WITH CROWNS AND CRYPTANDS . . . . . . . . . . I. INTRODUCTION. . . . . . . . . . . . . . . II. DETERMINATION AND INTERPRETATION OF THE LINESHAPES O O O 0 C C O O O O O O O O A. MEASUREMENTS IN THE ABSENCE OF EXCHANGE O O O I O O O C O O O O O B. EVALUATION OF EXCHANGE TIMES. . . . . III. RESULTS AND DISCUSSION. . . . . . . . . . A. LIGAND EFFECT ON EXCHANGE RATE. . . . B. SOLVENT EFFECT ON THE EXCHANGE PROCESSES O O O O O O O O I O O C O O C 0 DISCUSSION 0 O C C C C O C C I O O O 0 CHAPTER V. A CESIUM-133 NMR STUDY OF THE COMPLEXATION OF CESIUM TETRAPHENYLBORATE BY C222: DISCUSSION OF THE TYPES OF COWLEXES O I O C O C O O O O I O O O O O O O I 0 INTRODUCTION 0 O O O O O O O O O O I O O O Page 68 89 99 109 109 125 142 143 143 144 151 152 152 168 183 186 187 Chapter II. EVIDENI COMPLEZ CHAPTER VI. ST. FUTURE STUI I. L'nvaet II. scones: termites APPENDIX A. APPENDIX 3. APPENDIX c. APPENDIX D. PEFEREXCES , Chapter Page II. EVIDENCE FOR TWO TYPES OF CRYPTATE COMPLEXES . . . . . . . . . . . . . . . . 187 CHAPTER VI. SUMMARY AND SUGGESTIONS FOR FUTURE STUDIES. . . . . . . . . . . . . . . . 201 I. SUMMARY . . . . . . . . . . . . . . . . . 202 II. SUGGESTIONS FOR FUTURE STUDIES. . . . . . 205 APPENDICES APPENDIX A. . . . . . . . . . . . . . . . . . 208 APPENDIX B. . . . . . . . . . . . . . . . . . 212 APPENDIX C. . . . . . . . . . . . . . . . . . 221 APPENDIX D. . . . . . . . . . . . . . . . . . 235 REFERENCES. . . . . . . . . . . . . . . . . . . . 238 vi Table N l Thermodyne Reactions 2 Thermodan Cryptate by Calorie 3 Trends in crYPtands 4 Values of A. and thé Alkali Me‘ Values of TechfiiqUES The phYsi Nucleus . Physical 3 Correctior on DA-GO. cesium-l3 Salt SQIUI 10 LIST OF TABLES Table Page 1 Thermodynamic Quantities for Complexation Reactions of Macrocyclic Polyethers . . . . ll 2 Thermodynamic Quantities for Alkali Cryptate Complexation Reaction Measured by Calorimetry at 25°C in Water . . . . . . . 22 3 Trends in Thermodynamic Parameters with Cryptands . . . . . . . . . . . . . . . . . . 21 4 Values of the Average Excitation Energy, A, and the Expectation Value p for Alkali Metals . . . . . . . . . . . . . . . . 33 5 Values of qu Obtained by Various Techniques. . . . . . . . . . . . . . . . . . 37 6 The Physical Properties of the Cs-133 Nucleus . . . . . . . . . . . . . . . . . . . 38 7 Physical Properties of Solvents and Correction for Magnetic Susceptibility on DA—60. . . . . . . . . . . . . . . . . . . 55 8 Cesium-133 Chemical Shifts of Cesium Salt Solution at 25°C . . . . . . . . . . . . 60 9 Ion Pair Formation Constants of Cesium Salts in Various Nonaqueous Solvents. . . . . 67 10 Mole Ratio Study of 18C6 Complexes with Cesium Salts in Various Solvents by Cs-133 NMR at 25°C . . . . . . . . . . . . 72 vii Table 11 12 13 14 15 16 17 18 19 Hole Ratit CsTPB in J at 25°C . Mole Rati' CsTPB in 1! NMR at 25' Mole Ram CsTPB in x at 25°C . Ible Ratig CsTPB in I at 25°C . Formation with LigaI Mole Ratic with CSTpr ‘38°C . Limiting ( SOlVents . Half‘heig3 I 1 Va] + 0 CS ‘nganc L09 K Table 11 12 13 14 15 16 17 18 19 Mole Ratio Study of BBC Complexes with CsTPB in various Solvents by Cs-133 NMR at 25°C . . . . . . . . . . . . . . . . . Mole Ratio Study of DCC Complexes with CsTPB in Various Solvents by Cs-133 NMR at 25°C . . . . . . . . . . . . . . Mole Ratio of C2228 Complexes with CsTPB in Various Solvents by Cs-133 NMR at 25°C . . . . . . . . . . . . . . . . . Mole Ratio Study of C222 Complexes with CsTPB in Various Solvents by Cs-133 NMR at 25°C . . . . . . . . . . . . . . . . . Formation Constants of CsTPB [0.01 M] with Ligands in various Solvents (25°C) . Mole Ratio Study of 18C6 Complexes with CsTPB in Pyridine at 24°C, -6°C, -38°C . . . . . . . . . . . . . . . . . . Limiting Chemical Shifts in Various Solvents. . . . . . . . . . . . . . . . . Half-height Linewidth (Ava Hz) of Cs+-Ligand Complexes at 24°C. . . . . . . Log K Values of Cesium Salts with 1 Ligands in Various Solvents (25°C). . . . viii Page 75 77 79 81 88 93 97 101 108 Table 20 Concentr Shift at Complex 21 Data fro Referenc 22 hole rat Chemical 0f 18C6 : 23 The Simui and Least for the c borate by TemPeratu 24 Limiting Complex a for the p 1" Pyridi 25 Hole RatN ChemICal 26 Hole Rati ChemiCal the preSq 27 Hole Rati ChemiCal presenCe Table 20 21 22 23 24 25 26 27 Concentration Dependence of the Chemical Shift at 25°C for CsTPB and its 1:1 Complex with 18C6 in Pyridine . . . . . . . . Data from the Test of the Insulated Reference Tube. . . . . . . . . . . . . . . . Mole ratio-Temperature Data for the Chemical Shift of CsTPB in the Presence of 18C6 in Pyridine . . . . . . . . . . . . . The Simulated Formation Constant (K1) and Least-Squares Adjusted Constant (K2) for the Complexation of Cesium Tetraphenyl- borate by 18C6 in Pyridine at various Temperatures. . . . . . . . . . . . . . . . . Limiting Chemical Shift of the Cs+-(18C6)2 Complex and the Thermodynamic.Parameters for the Reaction Cs+-18C6 + 18C6 I Cs+-(18C6)2 in Pyridine at Various Temperatures . . . . . Mole Ratio-Temperature Data for the Chemical Shift of CsTPB [0.02 M] in the Presence of C222 in PC. . . . . . . . . . . . Mole Ratio-Temperature Data for the Chemical Shift of CsTPB [0.01 M] in the Presence of C222 in PC. . . . . . . . . . Mole Ratio-Temperature Data for the Chemical Shift of CsTPB [0.02 M] in the presence of C222 in DMF . . . . . . . . . . . ix Page 111 119 121 123 126 128 129 130 Table 28 N 30 31 32 M Mole Rati Chemical Presence Mole Rati Chemical Presence Formation Chemical 0f CsTPB Temperatt FOrmatior Shifts f< C222 in E “manor Chemical Cf CSTPB AcetOne a Table 28 29 30 31 32 33 34 Mole Ratio-Temperature Data for the Chemical Shift of CsTPB [0.01 M]in the Presence of C222 in DMF . . . . . . . . . Mole Ratio-Temperature Data for the Chemical Shift of CsTPB [0.02 M] in the Presence of C222 in Acetone . . . . . . . Formation Constants and Limiting Chemical shifts for the Complexation of CsTPB by C222 in PC at Various TemperatureS. . . . . . . . . . . . . . . Formation Constants and Limiting Chemical Shifts for the Complexation of CsTPB by C222 in DMF at Various Temperatures . . . Formation Constants and Limiting Chemical Shifts for the Complexation of CsTPB [0.02 M] by c222 in Acetone at Various Temperatures . . . . . Thermodynamic Parameters for the Complexation of CsTPB by C222 in Various Solvents. . . . . . . . . . . . . Temperature Dependence of the Transverse Relaxation Times for Free and Complexed Cesium Cations in Propylene Carbonate. . . . . . . . . . Page 131 132 137 138 139 140 147 Table 36 37 38 39 40 TemperatuI RelaxatioI in the Pre Temperatu: Time T, o Propylene Reciproca Time of c B). . . TemPeratu Time in PI the Corre RElaxaticI IB) . . . Exchange I Parameter Some Cesi proleene Tempe”tu RelaXatiC‘ Com918xed Temperatu Relaxatio (C222) CE) Table Page 35 36 37 38 39 40 Temperature Dependence of the Transverse Relaxation Times in Propylene Carbonate in the Presence and Absence of C222 . . . . . 148 Temperature Dependence of the Exchange Time I, of Some CsC+ Complexes in Propylene Carbonate and Corresponding Reciprocal Transverse Relaxation Time of Cs+ (Site A), and CsC+ (Site 8). . . . . . . . . . . . . . . . . . . . . . 161 Temperature Dependence of the Exchange Time in PC in the Presence of C222 and the Corresponding Reciprocal Transverse Relaxation Times of Cs+ (A) and Cs+C222 (8)162 Exchange Rates and Thermodynamic Parameters for Release of Cs+ from Some Cesium Macrocyclic Complexes in Propylene Carbonate . . . . . . . . . . . . . 166 Temperature Dependence of the Transverse Relaxation Times of Free and 18C6- Complexed Cesium Cation in Pyridine . . . . . 172 Temperature Dependence of the Transverse Relaxation Time of Free and Complexed (C222) Cesium Cation in DMF . . . . . . . . . 173 xi Table ll Temperatu Q 43 44 4S 46 47 Relaxatio Cesium Ca Temperatu Time in t and the ( Vers e R~e CsC+ (B) Temperat Time in and the Verse F Csc+ (r Tempere Time it the CO Relaxa Table 41 42 43 44 45 46 47 Temperature Dependence of the Transverse Relaxation Time of Free and Complexed (C222) Cesium Cation in Acetone. . . . . . . . Temperature Dependence of the Exchange Time in the Presence of 18C6 in Pyridine and the COrresponding Reciprocal Trans- verse Relaxation Times of Cs+ (A) and Csc+ (B). . . . . . . . . . . . . . . . Temperature Dependence of the Exchange Time in the Presence of C222 in Acetone and the Corresponding Reciprocal Trans- verse Relaxation Times of Cs+ (A) and Csc+ (8). . . . . . . . . . . . . . . . Temperature Dependence of the Exchange Time in the Presence of C222 in DMF and the Corresponding Reciprocal Transverse Relaxation Times of Cs+ (A) and CsC+ (8). Exchange Rates and Thermodynamic Param- eters of 18C6 Complex Exchange in Propylene Carbonate and Pyridine. . . . Test of the Dependence of the A82 Value onde................. Estimated Enthalpies of Formation of the Exclusive (A81) and Inclusive (A82) in Acetone, PC and DMF. . . . . . . . . xii Page 174 176 177 178 184 195 195 Table 48 Estimate Complexa 49 Activati Release Table Page 48 49 Estimated Entropy Values for the Complexation Reaction at 298°K. . . . . . . . 198 Activation Parameters for the Release of Cs+ from Cs+C222 in DMF. . . . . . 200 xiii Figure 10 11 12 Diben NaBr- CsNCS Crypt. Newly CIth; The F: near field centre 133CS t0 inI tempe] (C) 3: Paris) APpar; Prepai VESSe) vesSeI Block magnet Concer Chemit VarioK Figure UIthH 10 11 12 LIST OF FIGURES Dibenzo-lB-Crown-G (DBC) . . . . . . . Na8r-D8C-2H20 Crystal Structure. . . . CsNCS-tetramethyl-DBC (1:1). . . . . . Cryptand, C (k+1, mwl, n+1). . . . . . Newly Synthesized macrotricyclic cryptand . . . . . . . . . . . . . . . The Frequency of the Cs resonance near 28.013 MHz at fixed applied field as a function of molar con- centration in H2160, D2160, and H 133 18 2 o. Cs Shift data (CsCl'MeOH) relative to infinite dilution shifts at three temperatures (a) 298°K; (b) 308.3°K; (c) 326°K and (d) CsCl+H20 for com- parison. . . . . . . . . . . . . . . . Apparatus for storing solvent in the preparation of NMR sample. . . . . . . Vessel for solvent purification. . . . Vessel for DMF purification. . . . . . Block diagram of the multinuclear magnetic resonance spectrometer. . . . Concentration dependence of the Cs—133 chemical shifts of cesium salts in various solvents . . . . . . . . . . xiv Page 15 19 41 41 49 50 51 53 64 Figure 13 14 15 l6 17 18 19 20 Crowns Chemic functi TPB’] 0.01 N Chemic functi in var Chemic a func [CS+T [CSTP Chemi a fun [CS+T [CSTP Chemi a fun [CS+Tr [Csrpg N°nli shift Nonlir Figure Page 13 Crowns and [2]-cryptands . . . . . . . . . 71 14 Chemical shifts of Cesium-133 as a function of mole ratio of [l8C6]/[Cs+ TPB'] in various solvents. [CsTPB]T = 0.01 M . . . . . . . . . . . . . . . . . . 83 15 Chemical shifts of Cesium-133 as a function of mole ratio of [D8C]/[Cs+TP8'] in various solvents. [CsTPB]T = 0.01 M. . 84 16 Chemical shifts of cesium-133 as a function of mole ratio of [DCC]/ [Cs+TP8-] in various solvents. [CsTPB]T = 0.01 M. . . . . . . . . . . . . 85 17 Chemical shifts of cesium—133 as a function of mole ratio of [C2228]/ [Cs+TP8-]in various solvents. [CsTPB]T = 0.01 M. . . . . . . . . . . . . 86 18 Chemical shifts of cesium-133 as a function of mole ratio of [C2228]/ [Cs+TP8-] in various solvents [CsTP8]T = 0.01 M. . . . . . . . . . . . . 87 19 Nonlinear curve fitting of chemical shifts gs [18C6]/[Cs+TP8'] in DMSO . . . . 91 20 Nonlinear curve fitting of Chemical shifts Kg [18C6]/[Cs+TP8-]in MeCN. . . . . 92 XV Figure Page 21 A study of cesium-133 Chemical shift !§_mole ratio [18C6/Cs+] at 24°, -6°, -38°C in pyridine. [CsTPB]T = 0.01 M. . . 95 22 A typical fitting of a two step reaction . . . . . . . . . . . . . . . . . 96 23 The interaction of solvent molecules with complexed cesium ion. . . . . . . . . 102 24 Mole ratio - Cesium-133 chemical shift study for various ligands in DMF . . . . . 104 25 Cesium-133 chemical shift variation with the concentration of cesium salts in pyridine. . . . . . . . . . . . . . . . 112 26 Insulated NMR reference sample . . . . . . 116 27 Test of the insulated reference tube . . . . . . . . . . . . . . . . . . . 118 28 A three-dimensional plot of the cesium- 133 chemical shift gs mole ratio and temperature (°C) for solutions of CsTPB and 18C6 in pyridine. The concentration of CsTPB was 0.01 M. . . . . . . . . . . . 122 29 A comparison of simulated data (0 )‘with experimental data (A) for a chemical shift-mole ratio study at 297°K and 235°K for solutions of CsTPB and 18C6 in pyridine. . . . . . . . . . . . . . . . 124 xvi Figure 30 31 32 33 34 35 Cesiur ratio PIOPY tempe Cesiu ratio Prepy tempe Cesiu ratic at V6 A Plc tion aCet A ty 133 Figure 30 31 32 33 34 35 Page Cesium—133 chemical shift y§_mole ratio of [C222]/[Cs+TP8'] in propylene carbonate at various temperatures . . . . . . . . . . . . . . . 133 Cesium-133 Chemical shift 22 mole ratio of [C222]/[Cs+TP8-] in prOpylene carbonate at various temperatures . . . . . . . . . . . . . . . 134 Cesium-133 chemical shift !g_mole ratio of [C222]/[Cs+TP8-] in acetone at various temperatures. . . . . . . . . . 136 A plot of 2nK gs l/T for the complexa- tion reactions of Cs+ with C222 in acetone, PC, and DMF . . . . . . . . . . . 141 A typical KINFIT analysis of cesium— 133 lineshape for a solution containing 0.01 M CsTPB in propylene carbonate at -46°C. . . . . . . . . . . . . . . . . . . 146 Semilog plots of 1/T; for the cesium- 133 nucleus gs 1/T for solutions containing 0.02 M CsTPB in the presence and absence of various complexing agents. l/T; values 1 below about 15 sec- represent inhomogeneous broadening . . . . . . . . . . . . . . . . 149 xvii Figure 36 37 38 39 40 41 Semil | 133 T: tempe 0.02 PC.. Spect repr Ilef of e Spect a so] and I Spec. 80 and Spec Wit Figure 36 37 38 39 40 41 Page Semilog plots of 1/T; for cesium- 133 nucleus reciprocal absolute temperature for solutions containing 0.02 M CsTPB and 0.02 M Cs+C222 in PC.. . . . . . . . . . . . . . . . . . . . 150 Spectra at various temperatures for a solution containing 0.02 M CsTPB and 0.01 M C222 in PC. The dotted lines represent the chemical shifts of CsC+ (left) and Cs+ (right) in the absence of exchange. . . . . . . . . . . . . . . . 153 Spectra at various temperatures for a solution containing 0.02 M CsTPB and 0.01 M C2228 in PC . . . . . . . . . . 154 Spectra at various temperatures for a solution containing 0.02 M CsTPB and 0.01 M DCC (A) in PC . . . . . . . . . 155 Spectra at various temperatures for a solution containing 0.02 M CsTPB and 0.01 M 18C6 in PC. . . . . . . . . . . . . 156 Computer fit of the spectrum obtained with 0.02 M CsTPB and 0.01 M C222 in PC. The mixture of dispersion and absorption modes occurs because the first-ordered phase correction was not made with the spectrometer . . . . . . . . . . . . . . . 158 xviii Figure 42 43 44 45 46 47 Com} witl (A) Com; witl in I Arrh rate 1/T Arrh time Pros: SPCC' a so. 18C6 CoaL and . POOr chem: the e SPEC) a soI 0.01 Show in t. Figure 42 43 44 45 46 47 Page Computer fit of a spectrum obtained with 0.02 M CsTPB and 0.01 M DCC (A) in PC. . . . . . . . . . . . . . . . . 159 Computer fit of a spectra obtained with 0.02 M CsTPB and 0.01 M 18C6 in PC. . . . . . . . . . . . . . . . . . . 160 Arrhenius of plots of 1 (Exchange rates of Cs+ ion from ligands) gs 1/T for PC solution. . . . . . . . . . . . 163 Arrhenius plot of 1 (exchange time of Cs+ ion from C222) for propylene carbonate solutions. . . . . . . 164 Spectra at various temperatures for a solution of 0.01 M CsTPB and 0.005 M 18C6 in pyridine. The linewidth at coalescence were W400 Hz. This fact and the low concentration lead to the poor S/N. The dotted lines show the chemical shifts of CsC+ and Cs+ in the absence of exchange. . . . . . . . . . 169 Spectra at various temperatures for a solution containing 0.02 M CsTPB and 0.01 M C222 in DMF. The dotted lines show the chemical shifts of CsC+ and Cs+ in the absence of exchange . . . . . . . . 170 xix Figure 48 49 50 51 52 Spec 0.01 line CsC exch Semi nucl eith in p abOu‘ inhc Sem: 133 tai CSC 1/1 in' Ar CS br Figure 48 49 50 51 52 Page Spectra at various temperatures for a solution containing 0.02 M CsTPB and 0.01 M C222 in acetone. The dotted lines show the chemical shifts of CsC+ and Cs+ in the absence of exchange . . . . . . . . . . . . . . . . . 171 Semilog plots of l/T; for cesium-133 nucleus 23 l/T for solutions containing either 0.02 M CsTPB or 0.02 M Cs+-18C6 in pyridine. Values of l/T; less than 1 represent the effects of about 10 sec- inhomogeneous broadening . . . . . . . . . 175 * Semilog plots of l/T2 for the cesium- 133 nucleus 35 1/T for solutions con- taining either 0.02 M CsTPB or 0.02 M CsC222+ in DMF and in acetone. Values of 1 represent 1/T; less than about 10 sec- inhomogeneous broadening . . . . . . . . . 179 Arrhenius plots of 1 (exchange time for Cs+ ion from 18C6) in propylene car- bonate . . . . . . . . . . . . . . . . . . 180 Arrhenius plots of 1 (exchange time for Cs+ ion from the ligand C222) in DMF solution . . . . . . . . . . . . . . . 181 XX Figure 53 54 SS 56 Arrh for 1 solud A le 0f CS 15 te (dott ConVe SP801 Cs+cz low dOtt Figure 53 54 55 56 Page Arrhenius plot of r.(exchange time for Cs+ ion from C222) in acetone selution O O O O O O O O O O O O O O O O O 18 2 A plot of limiting chemical shift of Cs+C222 in DMF, PC and acetone XE temperature. The extrapolations (dotted lines) for all three solvents converge to the same value . . . . . . . . . 189 Spectra of solutions which contain the Cs+C222 complex and uncomplexed Cs+ at low temperatures in five solvents. The dotted lines indicate the peak positions of CsC222+ at room temperature . . . . . . 190 A comparison of simulated plot of limit- ing chemical shifts of CsC222+ in PC y§_temperature (°K) with experimental value where 0 means simulated value and x means experimental value . . . . . . 197 xxi CHAPTER I HISTORICAL not C1 becau: were : fact, and C tion teres devel POint in th thEir trans more Ethel SeSSt alka SYnt 1i9a Sand tion one “hit rea< We. 1. INTRODUCTION For many years, complexes of alkali metal ions were not considered an exciting area for chemical investigation, because it was assumed by most chemists that the complexes were neither stable nor important, and, as a matter of fact, alkali salts were frequently used to achieve high and constant ionic strength in solutions, when complexa- tion with other metal ions was occurring. Recently, in- terest in these alkali metal coordination complexes has developed, both from the chemical and the biological points of view, the latter because of their importance in the metabolism of plants and liver mitochondria and their significance as models for investigation of active transport processes in general. The first synthetic ligand, more or less specific for alkali catiors was a cyclic poly- ether obtained by Pedersen in 1967 (1). The ligand pos- sesses a bidimensional cavity which can accommodate an alkali cation. Soon thereafter Lehn and coworkers (2) synthesized cryptands which are diazapolyoxamacrocyclic ligands with tridimensional cavities. Both types of li- gand can form very stable alkali metal complexes in solu- tion as well as in crystalline form. This fact enables tone to more easily control and investigate the parameters which determine the characteristics of the complexation reaction. During the past decade, it has been found that alkali NMR spectra are very sensitive probes of the immediate envir mainl ion b nonaq II 1964, bioti rat 1 and s on mo indue Phore the s in ma macro promo Chara. in? S. M! Sized environment of the alkali metal ion (3-7). This work mainly concerns some aspects of the complexation of cesium ion by crown and cryptand complexing agents in various nonaqueous solvents as studied by Cs-133 NMR. II. HISTORICAL BACKGROUND Interest in the polyether type ligands has grown since 1964, when Pressman and Moore discovered that the anti- biotic valinomycin exhibits alkali cation specificity in rat liver mitochondria (8). Later, in 1966, Stefanaé and Simon (9) showed by electromotive force measurements on model membranes that alkali ion selectivity is mainly induced through complex formation of the antibiotic ion- phore with the cation in question. This observation was the starting point which stimulated scientific interest in macrocyclic complexation of metal ions. Synthetic macrocyclic polyether- crowns and cryptand ligands have very pronounced complexation abilities. Their special physical Characteristics will be reported separately in the follow- ing section. (A) MACROCYCLIC POLYETHER-CROWN COMPLEXATION OF ALKALI METAL IONS More than 50 macrocyclic "crown" ethers were synthe- sized by Pedersen, and many were found to solubilize alka of t numbt respe the I comp] alkali metal salts in non-polar solvents. The structure of the first crown compound, dibenzo-lB-crown-G (DBC) is shown in Figure l, where ”18" and "6" indicate the total number of atoms and the number of oxygen atoms in the ring respectively. Ultraviolet and infrared measurements of the D8C-KSCN system indicated the formation of a 1:1 complex (10). It was further shown that when the cation . / \._/°\__/ Figure 1. Dibenzo-lB-crown—G (88C). was too large to fit in the central hole of a cyclic polyether ligand, complexes with mole ratios of 1:2 or 2:3 (meta1:ether) could also be obtained (11). Some of the larger ethers have been shown to complex two metal ions simultaneously (12). Truter 35 31. (13) determined the crystal structures of several crown complexes. They showed that in the solid tt fO in Pec (ii) state, the alkali metal ion is located in the middle of the polyether ring and the ligands are in a gauche con- formation about the aliphatic carbon-carbon bonds. The interesting features revealed in their work and that of Pedersen are the following: (1) When the number of oxygen atoms is even and no larger than six, they are coplanar in the ring, and the apex of the C-O-C angle is centrally directed in the same plane. Symmetry is at a maximum when all the oxygen atoms are evenly spaced in a circle. When seven or more oxygen atoms are present in the polyether ring and the complexed cation is larger than the cavity diam- eter, the oxygen atoms cannot assume a coplanar configuration and, consequently, arrange them- selves around the surface of a right circular cylinder with the apices of the C-O-C angles pointed toward the center of the cylinder. (ii) The second interesting feature revealed was that, even in the crystalline state, there are inter- actions between the cation, anion and solvent molecules, as shown in Figure 2 for the NaBr- DBC-ZHZO complex (14). In this case, one sodium ion in ring A, is in a hexagonal bipyramid of oxygen atoms, six from the ligand and two axial.water molecules. The sodium ion in ring 8 is attached Figure 2. Na8r-D8C-2820 crystal structure. (iii) to a bromide ion, at one apex of a hexagonal bipyramid, the ligand forming the equator and water the other apex. The structure is held in a chain by water-bromide and water-water hydrogen bonds. Another example is the structure of the 1:1 compound CsNCS-tetramethyl-dibenzo-18C6 as shown in Figure 3. The compound shows equal sharing of the two thiocyanate ions between the two metal ions and a somewhat unsymmetrical en- vironment for cesium. It was also found that the stoichiometry in a unit cell may not necessarily be the stoichiometry of the complex. For instance, the unit cell of The “are MC Figure 3. CsNCS-tetramethyl-DBC (1:1). a 3:2 RbSCN complex (15) contains four molecules of a 1:1 RbSCN-08C complex and two uncomplexed polyether molecules of crystallization. The stability constant, K, of a complex is defined by K = [MCr+]/[M+JICr] where MCr+ is the complexed ion formed from metal ion (M+) and 1: Va stants techni select (18). goes t P01yet crown-l betwee] Sizes ; the Ca. and ligand (Cr) by the following reaction: K M+ + Cr 2 MCr+ (1.1) Values of the stability constants (or formation con- stants) have been measured by a calorimetric titration technique (16), by potentiometric measurements with ion- selective electrodes (17) and by spectroscopic methods (18). These studies revealed that the stability constant goes through a maximum for each cation with increasing polyether ring size. The maximum for Na+ is between 15— crown-S and lB-crown-6, for K+ is 18-crown-6: for Cs+ between 18-crown-6 and 21-crown-7. These optimum ring sizes are those which provide the closest fit between the cation and the ”hole". Pedersen and Frensdorff (19) noted that very few data are available on complexation reactions in solvents less polar than methanol. In solvents of lower polarity, ion- pair formation becomes significant so that the anion ef- fects would be appreciable. Smid gt 21. (18,20-22) in- vestigated the interactions of alkali metal ions and their fluorenyl ion pairs with crown ethers in tetrahydrofuran (THF) and tetrahydropyran (THP) by using Optical spec- trosCOpy, distribution equilibria, conductances, and viscosities. From their early conductance study (23,24) in THF, it is known that at room temperature, the salt exi CIOI 5011 only two plex pair where sents s°lver iOn pa dipole asSOci Consta: tic“, 1 metal j detail. talcum alkali exists predominantly as a tight ion pair but changes into a solvent-separated ion pair at lower temperatures. When crown ethers are added to fluorenyl salts in ethereal solvents, complex formation occurs. In systems where only 1:1 crown-ion pair complexes are formed, at least two isomeric complexes can be found, i.e., a crown com- plexed tight ion pair (FE,M+Cr) and a crown separated ion pair (FI,CrM+) (25). The equilibria in the solution are, F2“,M+ + Cr 2 F£-,M+Cr (1.2) -+ +_+ Fe ,M + Cr + F£,CrM (1.3) + + + ns + FI,M Cr + FI,CrM ,ns (1.4) where MCfIdenotes a complexed crown molecule, Fn' repre- .sents the fluorenyl carbanion and ns is the number of solvent molecules interacting with the crown-separated ion pair. A semi-theoretical calculation of the electric dipole moment was used by Grunwald (26) to obtain the association constant for ion-pairing. From the formation constant, K, one can obtain the free energy of complexa- tion, AGO. However, the entropy and enthalpy of alkali metal ion complexation have not been studied in muCh detail. Simon gt 31. (27-29) used a computerized micro- calorimeter to study the thermodynamic properties of alkali complexes of antibiotics, while Izatt gt_gl. (30-32) also to st synth 1 rev cycli tive. by Me tingt addit from IEplg is al Plexg (DCC) using of e) that equi: 10 also use a precision thermometric titration calorimeter to study the thermodynamics of formation of complexes of synthetic macrocyclic polyethers. The data shown in Table 1 reveal that, for the complexation reaction of macro- cyclic ligands, usually AHc and quite often, ASc are nega- tive. This is caused by the macrocyclic effect described by Margerum (37,38). He used this term in order to dis- tinguish it from the chelate effect because there is an additional enhancement in stability beyond that expected from the gain in translational entropy when chelates replace coordinated solvent from metal ions. Kinetics information about these complexation reactions is also very limited. Shchori gt_gl, monitored the com- plexation of the sodium ion by DBC and dicyclo-hexyl-18C6 (DCC) in dimethylformamide and methanol solutions (4) by using 23 Na NMR measurements. From a study of the variation of exchange rates with concentration, they postulated that the exchange mechanism involved the complexation equilibrium, Na+(X-), Crown 2 Na+(X-) + Crown (1.5) In both solvents, the activation energy is 12.610.6 Kcal/ mole for D8C and 8.3 Kcal/mole for DCC. The lower activa- tion energy for releasing Na+ ion from DCC was attributed to the flexibility of the macrocyclic ring. Recently the above authors also studied the decomplexation kinetics .THOSHDNHOQ OMHOMUOHOUZ MO WCOMUOUOZ COfiUUXNHQEOU MOM MOHUMUCUDO UwEfiCXEOEMOCE .H UNQSE 11 mm weoumoa u--- me.oa- mm mm Owe «.m- ee.mn m~ mm Omm m.e- eH.e- OH +s 14 mascuec mums -ofiaosoasxmeAa -e.m.~.a mm omzo a.e- m.m- mm +1 mm axe e- .e- mm +ez mums (canteen e.m.~.a mm ewe e.oH- o.m- mm +x mumauoheen (Hague: e.u.a mm Acmmxmomzv omxom m.a~- om.oa- mm mm Acumxmomzc omxoe H.H ee.m- mm mm Ao~mxmoozc eexee ~.o- ~m.m- mm mm lemmxmomzc ee\oe e.H Hm.~- mN mm lo~m\sowzc om\e~ m.e- m.H- mm mm lo~m\moozc e~\em e.ua- um.m- mm +1 mm Ao~m\momz. om\ee e.mu ~m.m- mm mm Acmm\mouz. oe\ee ~.m- me.m- mm mm Acmmxeomzc oexee m.m- me.~- mm mm lemmxeomzc emxem e.u- ee.e- mm +ez mumeuohemm ~.H . mom moo QHOE\HOO OHOE\HCOM 0o cofiemu ocsomEoo m< m< .osoe Oaaomoouomz .muonuomHom oaao>oouooz Mo mcoauooom cowusxoamaoo mom mowuwusmso Owenshoosuosa .H wanes I ~ n L>CONU 02. lfI I'll-ff IAN DODCWUCOU .u. 'HQIH 12 mm can e.HHu e.e- ea +nm mm ueoumoa u--- eo.ea- mm mm moo: m.esu m.e~- mu +s mm moan u--- e.~a-e mm mm Oman m.ma- e.e- mm mm cum m.H- mH.e- ee mm cam e.mu eo.m- mm mm o~m «.mau me.ma ea +1 mm mom: a.H- e.m- mm mm moan u--- e.u-e mm mm omzo v.0: m.~u mm oz m uoEOnH mm Owe ~.H- eH.~- mu mmz mm cum «.m- em.~u ee mm Omm e.mu He.~u ma mm o~m a.n- ee.~- OH +eo mm cum H.4a a~.nu as an ca: «.4- mm.m- mm mm ON: s.e- me.mu ea +nm am can e.~- em.mu ee . «om moo mace—\Huo macs—{HOOK o o CODCU 050980 m< m4 . mama. ofioaoouooz oosceueoo .H oases chVQDCfiUCOMV uH. TWA-MB 13 mm moms Hv.>+ hm.m- mm +M Camcocoz em some ma.m- o.e- mm +ez espouse: em moms ee.m+ me.- mm +ez mm moms He.ee- ~e.oa- me +s eeuoeeoz em moum HIOH x ma.m- ma.~t mm +M cflmmfiocwam> mm moms -- o.HH- mm +He mm mow: -- m.m- mm wmz mm some -- ~.He- mm +mo mm moms -- e.ua- mm +am mm moms -- m.HH mw +s mm mow: -- e- mm +ez onom-ONSmnAa -e.m.~.H mm cum m.o mo.- mu +84 mm cum m.e- He.m- mm wmz mm One e.e- om.m- 04 mm Omm m.m- em.m- ma .mmm woo oeoexaeo meoexaeus O. eonueo oeaoasoo ma m< .osoe Owaomoonooz .ooscwucou .H OHACB of I of e the for stil as - x-re Shoe very same to p less gate 1.8a 14 of DEC-K+ and DEC-Rb+ complexes in MeOH solutions by means of alkali metal NMR (5). The activation energy (Ea) for the decomplexation of the K+ ion is 12.6 Kcal/mole, while for Rb+, the exchange between free and complex sites was still indicated as being rapid even at temperatures as low as -50°. They interpreted these results as follows: The x-ray crystallograph study done by Truter gt_gl. (15), shows that both Na+ and K+ can fit the cavity size of BBC very well, therefore for these two ions they obtained the same Ea value. However, Rb+, a larger cation, is forced to protrude from the cavity plane and, consequently, is less tightly bound to the BBC molecule. The complexation reaction of dibenzo-30C10 with Na+, K+, Rb+, Cs+, NHI, Tl+ in methanol solution was investi- gated by Chock (39). The reaction mechanism he postulated is as follows: fast + + Cr k 2 + M+ % MCr+ (1.7) k21 Cr2 That is, a conformational transition was prOposed. The symbol Cr1 represents an unreactive species, Cr2 is an Open configuration which is ready to complex the cation, and MCr+ is a Closed configuration which is stabilized by a monovalent cation. His results also emphasized that the stability of the complex is dependent on the ionic radi Cuss DBC Thei is 2 thus Witt. 15 radius and the hydration energy of the cations under study. Cussler and coworkers (40) made conductance studies of BBC and DCC in acetonitrile (MeCN) and in methanol (MeOH). Their results indicated that the stability of the complex is affected by the solvation of the cation under study and thus solvent effects on the complexation reaction cannot be overlooked. (B) COMPLEXATION OF METAL IONS BY MACROHETEROCYCLIC LIGANDS-CRYPTANDS. Cryptands are polyaza and polyoxa macrocyclic compounds with tri-dimensional cavities (Figure 4). In order to Figure 4. Cryptand, C (k+1,m+1,n+l). accommodate different cations, the cavity size can be varied by changing the length of the ether bridge. The selectivity of complexation and stability of complexes, in general, are several orders of magnitude greater than for in tand For has firs ture Na PIOC than caus is i: WOrk, in e1 16 for crown compounds with the same number of oxygen atoms in the ring. In general, the term cryptand is ascribed to the free ligand and the term cryptate to the complex. The notation C(n+l,m+l,k+1) is an abbreviation for a cryp- tand with n+l,m+l,k+1 oxygen atoms in each branch. For instance, if a hexaoxadiamine macrocyclic compound has a value n=m=k=1, then it can be written as C222. The kinetics of the complexation by cryptands was first studied by Lehn 2E.El° (41) through a PMR tempera- ture study on the complexation of C222 with K+, Tl+, and Na+ in D20. They concluded that the exchange mechanism proceeded by a dissociation-complexation process rather than a bimolecular process. The symmetrical splitting caused by Tl+-H spin-spin coupling indicated that the ion is in the center of the molecular cavity. Dye and co- workers (42) studied the exchange rate of sodium cryptate in ethylenediamine. The activation free energy value they obtained is similar to that of Lehn obtained in aqueous solution. They also dissolved pure sodium in ethylamine (EA) and in THF in the presence of C222. Due to the com- plexation of the cryptand with Na+, the concentration of dissolved metal was greatly enhanced and a gold-colored (NaC222)+Na- salt was formed (433), At low temperatures (43b) they observed two NMR resonances, i;g;, (NaC222)+ and Na-, with the Na- peak shifted upfield about 63 ppm from saturated aqueous NaCl for both solvents, i.e., EA and THF. Kint com; the with The solu for C222 2911 by C- energ is re This 23 a] Studj and I 17 Kintzinger and Lehn (44) also used 23Na-NMR to study the complexation of Na+ ion with cryptands. They found that the 23 Na nuclear quadrupole coupling constant decreased with an increasing number of oxygen atoms in the ligand. The chemical shift (referred to a 0.25 M aqueous NaCl solution as external reference), had values of 11.15 ppm for Na+-C211,-4.25 ppm for Na+-C221, and -ll.45 ppm for Na+- C222. The line widths at halfheight were 13213, 4612 and 2911 Hz, respectively. Lithium-7 NMR kinetics studies have also been performed by Cahen, gt El° (6). They found that the activation energy for the decomplexation of the Li-cryptand complex is related to the Gutmann donor number of the solvent. This result contrasts with the kinetics studies of Shchori gt El! who used crown complexing agents. The latter authors studied the exchange rate of Na-DBC in DMF and methanol and noted that there seemed to be no solvent effect on the activation energy. However the donor number for DMF and methanol are very nearly same. In addition to the study of the complexation of metal ions with macrobicyclic ligands (denoted as [2]-cryptands), Lehn 25 31. also synthesized macrotricyclic ligands (de- noted as [3]-cryptands) (45,46). Later, they also studied the cation exchange rate between binding sites on two rings 130 NMR. Their inside the cavity of a [SJ-cryptate by observation for this study is summarized as follows: (1) Complexes display an intramolecular cation exchange proces is loc shown (2) '1 decree energk (3) z Preser 18 process which interconverts two species in which the cation is located unsymmetrically in the molecular cavity, as shown below. ‘Wfl \0/ .-.: ““0 v— ~~ 2.»0 (“Hi ‘ 33" (2) The free energies of activation, AG: for this process decrease with increasing size and decreasing hydration energy of the cations in the order Ca+2 > Sr+2 > Ba+2. (3) An intermolecular cation exchange process is also present, but its rate is much slower and its free energy of activation is much higher than those of the intramolecu- lar process. (4) Both intra and intermolecular cation exchange is fast for the weak complexes which form with alkali cations. Another [3]-cryptand has been synethsized recently (47). As shown in Figure 5, the attractive feature of this molecule is that it possesses a spherical intra- molecular cavity into which the cation may be placed. Preliminary measurements show that the logarithm of the stability constants for the K+, Rb+, and Cs+ complexes in water are about 3.4, 4.2 and 3.4, respectively; and Fig the] Pera‘ (at. cryp Obta flow kins mure ca+2 exo- and fiQU; the 1 has . 19 no Q 521} (We?) UV Figure 5. Newly synthesized macrotricyclic cryptand. the PMR kinetic study obtained from the coalescence tem- perature showed that the free energies of activation for K+, Rb+ and Cs+ are 15.5 (at 28°), 16.7 (51°) and 16.1 (at 41°C) Kcal/mole. Kinetics information about the complexation between cryptands and metal ions is scarce and has been mainly obtained by NMR techniques. Very recently, a stopped- ‘flow technique (48) was used for a kinetics study by Wil- kins g£_al. They followed the color variation of the +2 *2 murexide-Ca and +2 complexation reaction of murexide-Ca Ca -cryptate. In this study, conformational changes, i;g;, exo-exo endo-endo (41,49), were also considered. Endo and exo configurations are shown below. The endo con- figuration has the lone pair electrons directed toward the interior of the cavity while the exo configuration has the lone pair electrons turning outside. Formation constants of alkali metal cryptates in aque SiVe niqt tior an a com; a pc C22: reSt ing This the Of p Valu the from Stil 20 exo-exo endo-endo exo-endo aqueous and methanolic solutions have been studied exten- sively by Lehn and coworkers who used potentiometric tech- niques. Recent reviews (50,33) contain extensive compila- tions of complexation constants. Cahen 22.2l- (7) used an alkali NMR technique to study formation constants of complexation in several solvents. They found that, in a poor donor solvent such as nitromethane, the addition of C222 or C221 cryptand to a lithium perchlorate solution resulted in a drastic chemical shift which reached a limit- ing value at higher than 1:1 (ligand/metal) mole ratios. This phenomenon indicates that the formation constant of the resulting cryptate is large in contrast to the case of pyridine, a better donor solvent, in which the limiting value is not reached even at a 25:1 mole ratio because of the competitive action of the solvent. The thermodynamic quantity AGO can be obtained directly from the formation constant, but AH0 and AS0 values are still very limited in extent. Only very recently Lehn and an a “YE Tabi com; char The: and cal< Sr+j Na+ //“'1 [Ca' hav‘ S‘flm 21 and coworkers (51) made calorimetric measurements under an argon atmosphere to avoid the problems caused when cryptand solutions absorb C02. Their data are listed in Table 2. In this work, they stressed explicitly that the complexation reactions to form cryptates have large negative changes in enthalpy, sometimes, negative entropy changes. These are important factors in determining the stability and selectivity of the complexation reaction. From their calorimetric study they observed the trends shown in Table 3. Table 3. Trends in Thermodynamic Parameters with Cryptands. Cation Complex With Cryptand Dominant Minor +2 +2 . + Sr ,Ba (not With C222),Na (C222) AH<0 TAS>O Na+,K+,Rb+,Cs+ AHO AH<0 . ' .+ +2 In addition, they noted that [L1 CC ], [Ca CC ], 221 211 +2 [Ca CZC222] are entirely entropy stabilized with about zero heat of reaction. Crystal structures of a number of alkali cryptates have been determined by Weiss and coworkers (52-55). For [2]-cryptates it was observed that the metal ion is centro- symmetrically located in the cavity. In some cases, solvent .HOUQK CH OomN U6 NHuTEfiHONQU >n~ mumwhfinflmuz COfiUOUGK COfiUflXTHQEOU @uuuflxnhUlfiHMXHaw MO WTHUflUEQDU UsEfiCkmuOEMWCL. .N @HQQE 22 .cofiusuoan mouocop s .musum omxoameoo o» Hommcsnu monocoo o .UMHAUG+ZH + UuAAV + Amsscc+z smusnssa saucmumsufis was mo coaussuom spas sssmsa swusussa was oucw mmsnm msm on» Eoum coauso on» no Hmmmcsuu ..o.w .acowusmflq: movocmv a muonz nms + oms u «ms ans + ems u ass "maoz ~.o~ h.a~ m.ma 1111111111 am<| Anw.mv H.mh m.vm «.mw llllllllll «ms: «mm an Inn: Hm mm 111:: amqu Afim.mv Hm lulu hm maa lllll «ms: «mm In: m.ma h.am m.ma m.HN am<| A«N.~V nan m.vm m.~m m.HHH H.~ma amsn Ham :1: In: In: m.mm mm «ms: Ano.av In: 1|: 1|: ~.HHH ~.>ma «as: Haw laws.ac+mo Asms.av+nm Asmm.av+m Asmm.oc+sz Assn.oc+fiq meos\Hsoxv sass“; o o o o o . nouoEsHsmosuosa .uousz ca comm us muuosauoaso >n omHSmsoz cowuoswm cowusxoameoo manuaauolaasxa< mo mowuwucsso oafiscmoosuona .m magma 23 molecules are also linked to the complex, e.g., Rb+(C222) 2 +2 sew-H20, 13:1+ (c222) (somzazo (56) and Ba (c322) (somz- 2820 (57). The solvent molecule or anion can extend into the ligand cage and coordinate with the trapped metal ion in the center. It was also found that for [2]-cryptates the preferred configuration is endo-endo in the crystal. For [3]-cryptates (58,59) the metal ion is also located in the cavity. Monovalent ions usually form weaker complexes with [3]-cryptands but 1:2 (ligand/metal) complexes can be formed. The crystal structure of an Ag+-[3]-cryptate complex actually shows two silver ions located in the two rings and linked with a third Ag+ through the oxygen of the nitrate group of AgNO3. This third Ag+ is outside of the cage. In summary, the complexation reactions of metal ions with macrocyclic ligands depend on the following factors: (1) The type of binding site in the ring. For a ligand with a donor atom such as 0, N or S, the stability usually follows the trend, 0 > N > S for small ions such as Li+, Na+. For a large ion such as NH: or a group B ion such as Ag+ the same trend may not necessarily hold because of some covalency in the coordination. (2) The number of binding sites in the ring. Compare for example the log R value for the Na+ and K+ complexes of C222 (with Na+ or K+ in MeOH log K is larger than 8) and 'czzc8 (with Na+ in MeOH log K is 3.5 while for K+ it “Y the sit of (3) (60 she the sig sut dex his Cai th: th. Qf; is 11. 24 it is 5.2). In this comparison, C22C8 denotes a [2]- cryptand with one aliphatic branch. Both ligands have about the same cavity radius but C22C8 contains two fewer oxygen sites. This results in a decrease in stability by a factor of 104-105 (50). (3) The relative sizes of the metal ions and the ligand (60). The K value as a function of the cation radius shows a maximum for any given ligand. This indicates that the ratio of cavity radius to the ionic radius contributes significantly to the selectivity of these ligands. (4) Steric hinderance and ligand thickness (61). The substituents on the ring will introduce rigidity and hin- derance in the ligand. These can make the ligand have a higher selectivity and ability to discriminate against cations which are either smaller or larger than the preferred one. The ligand interposes a layer between the cation and the outside medium. Therefore the thicker the ligand, the better the cation is shielded from the medium. This effect decreases long range ionic stabilization. The effect is larger the higher the dielectric constant of the solvent. (5) The solvent and extent of solvation of the ion and ligand. For example, Cahen's study illustrates that in solution the complexation reaction competes with solvation. Media which can solvate cations more strongly usually result in weaker complexes. (6) The electrical charge of the ion. Previous studies 25 have shown that, with a similar cation radius, a bivalent ion forms a stronger complex than a monovalent ion (6)). However, in this case ligand thickness becomes very im- portant because the long range interaction energy varies as the square of the ionic charge. (7) Topology of the ligand. The dimensions and geometry of the ring can greatly affect the stability of the complex. This is most pronounced when the complexing groups are held in favorable positions by the ligand framework. The extent of ion-pairing also depends on the geometry of the ligand. In addition to the recent work of Smid gt 31,,electron spin resonance studies (62,64) have been made of alkali metal hyperfine splittings in the presence of macrocyclic poly-i ether ligands. These results illustrate that several types of ion-pairs are formed with crown complexes in low dielectric media. However this effect was not observed when cryptates were in the same media. (C) NUCLEAR MAGNETIC RESONANCE ( i ) INTRODUCTION Electrolyte solutions are particularly suited to in- vestigation by nuclear magnetic resonance techniques. The 7 presence of extremely rapid and generally random molecular motions averages local magnetic and electric fields to very small values and can result in narrow resonance lines even for quadrupolar nuclei. This fact is important bec and In Che eff in in in In th at of ar. si tC m 1‘4 m.- i} W; fl 26 because it enables small differences in magnetic shielding and/or fine structure of the resonance signal to be detected. In addition, broadening of the resonance line because of chemical exchange, hyperfine interaction or quadrupolar effects may be studied and chemical information derived from the observed behavior. Proton NMR is useful in the investigation of solvent or ligand behavior and can provide information about ion-solvent or ion-ligand interactions. In favorable cases, non-proton NMR can be used to study the ions themselves and thus provide direct information about such interactions. Although resonance frequencies of metal ions are sensitive to ion-solvent, ion-ligand, and inter-ionic interactions, the generally weak resonance signals and the special instrumentation required combined to make such studies rare prior to the last decade. All alkali metals possess at least one isotope with a magnetic nucleus; e. ., 7Li, 23Na, 39K, 87Rb, and 13303. The in- herent intensity of the resonance is much lower than that of the proton. However, the sensitivity of the nuclear magnetic shielding constants to the nature of the surround- ings is considerably larger and increases as the atomic number of the ion increases. This condition leads to a wide range of chemical shifts, from a few ppm to hundreds of ppm. Variations in the chemical shift result primarily from changes in the paramagnetic shielding constant, op. spi the in and tex Var at ma) whe 27 (ii) CHEMICAL SHIFT STUDIES OF ELECTROLYTE SOLUTIONS For an assembly of identical but isolated nuclei of spin I and magnetic moment u in a static magnetic field Ho, the simplest expression for the resonance condition is v0 = uHo/Ih = yHo/Zn (1.8) in which v is the frequency at which energy is absorbed o and y is the magnetogyric ratio. The latter has a charac- teristic value for each isotopic species. However, a variety of mechanisms may produce secondary magnetic fields at a nucleus. The actual field experienced by the nucleus may be written as H = Ho (1-0) (1.9) where o is a dimensionless quantity known as the shielding (or screening) constant. Ramsey (65,66) has developed general theoretical expres- sions for chemical shifts caused by magnetic shielding of nuclei in molecules. This treatment has been applied succes- fully to simple molecules but the approximations required to apply it to more complex systems yield only qualitatively cor- rect results. Saika and Slichter (67) attempted to explain the difference in the shielding constants of F2 and F-. They divided the contributions into separate terms: (1) the diam a pa cont expa envi be; have ider Rams ind; sone Fret The) Mutt 28 diamagnetic contribution for the atom in question. (2) a paramagnetic contribution for the same atom and (3) contribution from the electrons of other atoms. When investigating an electrolyte solution, one can expand Ramsey's formulation as applied to a solid state environment. If alkali halide crystals are considered to be purely ionic in character, each constituent ion will have a spherical closed-shell electronic configuration identical to that expected for the isolated ion. In Ramsey's expression, the shielding constant of the ionic nucleus is determined only by the diamagnetic term, ad, for an isolated ion. Therefore, the observed large.para- magnetic contributions found for crystals (usually 102- 103 times larger than ad) indicate that there are additional interactions present which are able to distort the symmetry of the electron distribution and introduce some net orbital angular momentum into the ion. This perturbation gives rise to a paramagnetic chemical shift. Such inter- actions might be considered in terms of electrostatic, covalency or overlap effects. However, the calculations indicate (68) that neither electrostatic effects nor rea- sonable estimates of the degree of covalency can account for shifts of the magnitude observed. The best inter— pretation so far is that of Kondo and Yamashita (69). They suggested that the cause of the paramagnetic shift is due to the short range repulsive forces between ions. Mutual overlap of atomic wave functions of neighboring i0: Pa ra to 9C. fc t} ac La L, h 29 ions produces a strong repulsive force mainly due to the Pauli exclusion principle. This force acts over a short range and competes with the electrostatic forces which tend to reduce the separation of oppositely charged ions. At equilibrium in the crystal the attractive and repulsive forces are in balance. Kondo and Yamashita used L6wdin's (70) orthogonalized-atomic-orbital model and considered that the overlap integrals are significant only for inter- actions of orbitals which belong to nearest-neighbor ions. Later, both Ikenberry and Das (71) Hafemeister and Flygare (72) gave more exact derivations of the paramagnetic shift to be expected from overlap forces in alkali halide crystals. But at that time they ran into difficulty in comparing theoretical results with experiments because the experimental shielding data are usually referred to a reference sample in which the shielding constants are not known. Therefore it was thought that this problem could be solved by referring all experimental shielding constants to the infinite dilution chemical shift in water which, in principle, is supposed to be constant because of the strong hydration properties of water. When the alkali salt is put into a solvent, the cal- culated paramagnetic shielding constant, op, from the crystal state and the experimentally measured shift, 6, referred to the aqueous solution are related by the equation: ar 0‘ p] mc on si re Va in ac fr Ci Di 5‘! re 90: 30 o = o - 6 (1.10) where gig is the paramagnetic shift of the hydrated ion 0 aq of the partner ion in any alkali metal or halide series, relative to the "free ion". Since a should be independent 0 aq obtained by a combination of theory the constancy of o and experiment, therefore, provides a test for the proposed overlap mechanism and the accuracy of the wave functions em- 0 aq is not constant. Attempts were made to explain the dis— ployed in the calculation. However, in reality, a crepancy by Hafemeister and Flygare (73), Ikenberry and Das (74) and Y. Yamagata (68) individually by using various models in the calculations. However, all of them could only illustrate that ion-solvent interactions can produce sizable paramagnetic shifts of the ionic nucleus, but their results were still not comparable with the experimental values. Therefore, it was suggested that chemical shifts in solutions may be caused not only by ion-solvent inter- actions in dilute solution, but also may have contributions from direct interionic effects. By using the Rondo-Yamashita model one could in prin- ciple account for the chemical shifts in aqueous solutions. Direct collisions between ions will distort the spherical symmetry of the electron distribution and also can produce paramagnetic chemical shifts. In the solid state, the relative positions and distances of separation of the component ions are known, and, therefore, the theoretical ur se ma sh fo bo wa th. se] ma] the fir the Pa: “he 31 attack is easier. In solution the environment of each ion varies randomly with time as the ion and solvent molecules undergo rotational and translational diffusion. The ob- served chemical shift is an average value resulting from many separate contributions corresponding to the various short-lived associations in solution. An exact expression for the shielding constant would require a knowledge of both the radial distribution function of other ions and water molecules about the central ion and the magnitude of the appropriate overlap integrals as a function of the separation. This information is difficult to obtain. To make this problem tractible, Deverell gt El! (75) modified the Rondo-Yamashita theory in the following way: At in- finite dilution, the only interactions present are between the ion and the water molecules. This contribution to the paramagnetic chemical shift can be expressed by 0 -16a2 -3 o an =‘—_K_— (ri >p Ai-HZO (I'll) where A is the average excitation energy, a is the fine-structure constant, p is the expectation value of r;3 for an outer p- electron of the central ion, i. Ai-H O is an appropriate sum of the squares of overlap 2 integrals between the outer p-orbitals of the ions, i, and the outermost orbitals of neighboring water molecules. The superscript denotes zero the The wher tegr all ion The and 501v Prob and the: p[ZAi-j + (AC -A° )] (1.12) where Ai-j is the sum of the squares of the overlap in- tegrals for the interionic contribution and is taken over all ions in solution other than the central ion, 1. The ion j may be of like or unlike charge with respect to i. The superscript c denotes the concentration of the salt, c o . and (Ai_H20 Ai-HZO) represents changes in the effect of solvent-ion overlap integrals. c c Both terms, Ai-j and Ai-HZO' are dependent upon the probabilities of collision occurring between the central ion and other ions or water molecules in solution and are therefore concentration dependent. Also, from the factor -3 i for the observation that the magnitude of the chemical /A, as shown in Table 4, it is possible to account shift increases with the atomic number of the alkali ion. The following is a recapitulation of the previously described development of alkali metal ion chemical shift studies: Ramsey used standard perturbation theory to express the shielding constant as (67): Tab] II I» 33 Table 4. Values of the Average Excitation Energy A, and the Expectation Value for Alkali Metals.a p -3 -3 (1:1 >p A (1:1 > Ion a.u. Rydbergs A + Na 16 2.72 5.9 K+ 12.94 1.62 7.98 Rb+ 20.22 1.47 13.8 Cs+ 23.42 1.25 18.7 aReference 76. Si no ti EX it tc CI t1 34 o = Gd + op 2A A A 2 r 2 - r r = 62 {(lpolék 3 kkl wo>+ 2mc rk -1 A Ek {(so-Em) [(Wollszl‘pmxi’mlg "3W0” etc]} (1.13) m r k Since excited state wavefunctions and energies are usually not known, Ramsey introduced the average energy approxima- tion and the paramagnetic shielding constant was then expressed as follows: 0 - (- e2) (I 14) _ _____ ' , P AM2C2 o ké' k k k 0 A :average excitation energy. e :electronic charge M :electron mass. C :velocity of light. p) :angular momentum of the k'th electron, and rk :radial distance of the k'th electron from the origin at the nucleus. The subscript 0 refers to the ground state and m refers to the excited state. For a crystal, an additional interaction force was introduced. Two models have been proposed in an attempt to rationalize the large paramagnetic shifts, op, of ionic crystals. These are Yosida and Moriya's (77) charge- transfer covalency model (YM model), and Kondo and Yamashita's (69) over of the Qt halide m plausibl derivati transfer 18p betwe ion with introduce in Ramse} integral 1' and ti Their cori Pulsive 1 By Combix mOre C10: from 51k.- also uSe. the KY m. atomic w but also Their Ca Can be (next he ”911 as P01. relate w 35 (69) overlapping-ion model (KY model). Later, calculation of the quadrupole coupling constants in diatomic alkali- halide molecules (78) showed the KY model to be more plausible. However, Das gt_gl. (71) gave a more exact derivation by including the effects of overlap and charge transfer covalency. For the KY model they considered over- lap between the outmost s and p orbitals of the central ion with those of neighboring ions. Das gt El: not only introduced p for a p electron of the central ion, i, in Ramsey's expression, but also included the overlap integral between the outer p—orbitals of the central ion, i, and the outmost s and p orbitals of all other ions, j. Their conclusion was that the KY model (short-range re- pulsive force) is predominant over the YM model (covalency). By combining the two effects, the resultant calculation more closely approximates the experimental data obtained from alkali halide crystals. Hafemeister and Flygare (72) also used the symmetrical orthogonalization method as did the KY model, but they not only considered the overlap of atomic wavefunctions of nearest-neighbors in the lattice but also included the next nearest-neighbor's interaction. Their calculations indicate that alkali-alkali interactions can be neglected but for the halide ions, halide-halide (next nearest-neighbor interactions) should be included as well as halide-alkali interactions. For the case of solutions, however, op still does not relate well to the observed values. Yamagata (68) considered the dipc molecule shifts < ions bui computec approacl meantime Hartree- Give the Stantia] and cowc Cept to eXpresSj describe be made about t} when Lut results. optiCall obtaine pariSOn Shifts 1 given in 36 the dipole polarization of the ion by adjacent water molecules. Calculations based on this mechanism predict shifts of the correct order of magnitude for the halide ions but not for alkali metal ions. Ikenberry and Das computed c§2+ for a model of Rb+(H20)6 by using the same approach as they did for alkali-halide crystals. In the meantime, Hafemeister and Flygare also used D. Mayer's Hartree—Fock wavefunction to calculate 022+. Both results give the right order of magnitude but still disagree sub- stantially with the experimental observations. Deverall and coworkers applied a short-range repulsion force con- cept to solutions, as shown in Equation (1.12). With this expression, the concentration-dependent phenomena can be described but complete calculation of the shifts could not be made at the time because of the lack of information about the experimental value of egg, This problem was solved 0 when Lutz (79) determined 0a directly by combining the results obtained from precision NMR, atomic beam, and optical pumping experiments. A summary of 0: values obtained by different means is listed in Table 5. A com- parison of experimental results with calculated chemical shifts in halide crystals and aqueous solution has been given in Ceraso's thesis (95), Page 39. Um. . mmwSmufiCSUmwB TDOflHflxw \AQ VGCHQUQO 00 N0 mums-N85 M THQUE 37 me «use x a.u- passeuomxm coeusaom +nm muse oouuomou macauosuoucfl osasb oz coe1s0fi Hoowmcoo mu van msHHo>o soflusaom +bm moussoflm was Haouo>oo ms «use x eo.ou :oeusueusaom seamen Acmmv+am assumes» osxsw cofiuocsum>s3 xoomnoouuusm no vuoa x «.0: a.uoasz .m .0 coavoaom +bm snowman was Houmonomsm vb sled x mm.o1 mmauo>o macamv+nm use was humoncoxH .mom coH ooum m> Emacssooz H0002 Honusm woo +2 mmsvwsnooa msowus> ha omswmuao Us on no mosas> .m manna (iii) The in Tabl Table 6 The was fir: authors with Cor large 6} cesium i adding ‘ linearl. the Shi 38 (iii) CESIUM NUCLEAR MAGNETIC RESONANCE The physical properties of the Cs-133 nucleus are shown in Table 6. Table 6. The Physical Properties of the Cs-l33 Nucleus. Resonance frequence in MHz for a 1.4 T. field 7.87 Natural abundance, % 100 Relative sensitivity (vs. 1H) for an equal number of nuclei _2 at constant field. 4.74 x 10 Magnetic moment in multiples of the nuclear magneton (eh/4n mc) 2.5642 Spin 1, in multiples of h/2n 7/2 Electric quadrupole moment Q, in multiples of barns2 0.003 The magnetic resonance of cesium ions in solutions was first studied by Gutowsky and McGarrey (80). The authors observed that the 133C8 resonance frequency varied with concentration. Carrington 22.21: (81) also observed large shifts of the cesium resonance upon variation of the cesium halide concentration. They studied the effects of adding various salts to CsCl solutions. The shift varied linearly with the mole fraction of added salt and in general the shift was greater the larger the anion. rise to 39 More recently Deverall and Richards (75) studied the chemical shifts of alkali halide and nitrate aqueous solutions by the alkali NMR technique. The magnitudes of the shifts increased considerably with increasing atomic number of the cation, and shifts to both higher and lower fields relative to the cation at infinite dilu- tion were observed. The resonances of K+, Rb+ and Cs+ were observed to vary linearly with the mean activity of the salt and, for all alkali cations, anions showed a definite series of Shielding effects, geee, the order of increasing shielding was 1' < Br- < Cl' < F' < H20 < N03. The authors concluded that the chemical shifts of the cation resonances were not only caused by changes in the inter- actions with solvent molecules, but also with counter- ions. By a modification of the theory of Kondo and Yamashita, and the expression of Das EE.El-v the chemical shift at infinite dilution was formulated as Eq. 1.11. The concentration dependence of the chemical shift may result from two processes. Firstly, the addition of other ions may modify the ion-solvent interactions which give rise to an' and, secondly, direct interactions between the ions during collisions may also contribute to the shift. So the chemical shift at concentration C, relative to the free ion can be written as 2 1 l C C ] O = -16a '<_3’>np A Aion-ion + Aion-water (1'15) 40 where Ac is a sum of the squares of overlap integrals at concentration of salt C. Deverall eg_gl.(77) also found that the approach of the two oppositely charged ions will be activity dependent, and the term Igon-ion in the above equation should increase linearly with the activity of the solution. The increased paramagnetic shielding caused by halide ions of high atomic number is mainly caused by the higher values of the overlap integrals for the larger anions. The small shifts observed for Li+ and Na+ with concentration in aqueous solution are probably caused by strong hydration which inhibits direct cation-anion inter- actions. The increase in the interaction from K+ to Cs+ may be partly due to a decreasing strength of hydration and greater facility of approach of the anion to the larger cations, as well as to the increase in size and greater degree of overlap. Halliday and coworkers (82) experimented with isotopic solvents and observed Cs+ resonance shifts in dilute salt solutions. They studied the Cs+ resonances, at fixed applied field and as a function of molar concentration, 2160, D2160 and H2180. The results are shown in 2160 and H2180 are indistinguishable and extrapolate to the same value at in H Figure 6. The shifts observed in H infinite dilution. However, the shift at infinite dilu- tion in n215 that the isotope effects cannot simply be a function of 0 lies at a lower value. These results showed the masses of the molecules and suggest strongly that at Figure 5 Figure 7 41 "32Ur . / L c: at 28 MHz , * / / I) , ./ : // v’ / -4OU N I . Hfo 0 H,'°o o 02:60 b /o O - 500/ 0 on 04 06 C (Molar) Figure 6. The frequency of the Cs resonance near 28.013 MHz at fixed applied field as a function of molar (82) concentration in H2150, D2150, and H2130. // 100 Shin/Hz l L l l 1 J .05 leCHlmol/dm’l l l J Figure 7. 133Cs shift data (CsCl-MeOH) relative to infinite dilute shifts at three temperatures (a) 298°K; (b) 308.3°K; (c) 326°K (84) and (d) CsCl + H20 (83) for comparison. infinitii in the s from Hzi ance cat from “ox somewhat Wlth Hzc to cause At PUblishe solvents They ft: their 0‘ of 0.03 which a gives a the Val aqueOus F01- the Cs-l33 42 infinite dilution, the isotOpe effect is caused by a change in the strength of the solvent hydrogen bonds in passing from H20 to 020. The paramagnetic shift of the Cs+ reson- ance caused by adjacent water molecules arises mainly from "overlap" interaction of the water molecules. The somewhat stronger hydrogen bonds formed by D20, compared with H20, might easily modify the paramagnetic shift so as to cause the observed diamagnetic shift. At about the same time, Sharp and coworkers (83,84) published an alkali metal NMR study of Cs+ in various solvents at lower concentrations and at various temperatures. They first studied cesium salts in aqueous solution and their observations indicated that in the concentration range of 0.03 to 2 M, the data follow an empirical relation in which a plot of log (shift) 23. log (molar concentration) gives a straight line with slope 9. Later, they also tested the validity of this empirical logarithmic law for non- aqueous solutions and found it was much less satisfactory. For the investigation of the temperature dependence of the Cs-l33 chemical shift, they studied the systems CsCl-MeOH and CsBr-Hzo at a series of different concentrations. The shifts varied only slightly with temperature as shown in Figure 7. It should be noted that the reported chemical shift value at various temperature was obtained by first referring to a 2 M CsCl aqueous solution as an external reference measured at the same temperature and then further related to the value at infinite dilution at that temperature. III. Ct Fro of alkar interes tions. tion re: nonaque IS a $8 SOlutio StUdy o Prepert IV) of ‘ tands 14 43 III. CONCLUSIONS From the above discussion, it is evident that a study of alkali complexes with macrocyclic ligands is chemically interesting. In addition, it should have many useful applica- tions. Information on the thermodynamics of these complexa- tion reactions is very sparse, especially for reactions in nonaqueous solvents. It is also evident that alkali NMR is a sensitive probe of the complexation of alkali ions in solution and therefore is a very useful technique for the study of cryptate complexes. This thesis reports an in- vestigation by the Cs-133 NMR technique of the equilibrium properties (Chapter III) and dynamic properties (Chapter IV) of the complexation of cesium ion by crowns and cryp- tands in various nonaqueous solutions. CHAPTER II EXPERIMENTAL PART 44 I. SYNTHESIS OF CESIUM TETRAPHENYLBORATE AND LIGAND PURI- FICATION. Cesium tetraphenylborate (CsTPB) was made by mixing a tetrahydrofuran (THF) (Burdick and Jackson Laboratories Inc.) solution of sodium tetraphenylborate (NaTPB) with a concentrated aqueous solution of cesium chloride (Ventron Alfa Products, 99.9% pure). The fine white precipitate was washed continuously with conductance water: sodium contamination was checked with a flame emission spectrom- eter (EU-703) and was found to be less than 0.01%. Cesium triiodide solutions were made by mixing equimolar amounts of cesium iodide (Ventron Alfa Products, 99.9% pure) and iodine (Baker and Adamson) in an appropriate solvent. Crown ligand 18C6 (PCR, Inc.) was purified by forming a complex with acetonitrile (Mallinckrodt A. R. grade). When about 50 grams of 18C6 was dissolved at ambient tem- perature, in 125 m1 of acetonitrile (MeCN), fine white crystals of the 18C6°MeCN complex were formed. The flask was cooled in an ice-acetone bath to precipitate as much complex as possible and the solid was then collected by rapid filtration. The hygroscopic crystals were trans- ferred to a round-bottom flask equipped with a magnetic stirring bar and a vacuum take-off. The weakly bound MeCN was removed from the complex by pumping under vacuum for a few hours. The m.p. of recrystallized 18C6 was 39°C, the same as that reported (85). 45 Cry Inc. an hexane I of the with 8 Cry was use chased and 8), these t used f0 rEports 46 Cryptand C222 was obtained from E. M. Laboratories, Inc. and was purified by two recrystallizations from n- hexane followed by vacuum sublimation. The melting point of the snow-white sublimate was 68° (reported 68°C (86)) with a melting range of less than one degree. Cryptand C222B was a gift from E. M. Laboratories and was used as received. Dicyclohexyl-18C6 (DCC) was pur- chased from duPont Co. as a mixing of the two isomers (A and B). For lineshape analysis purposes, a separation of these two isomers was carried out, and only isomer A was used for the study. The method used was based on two reports (87,88). Approximately 100 g of Alumina Absorption (Fisher) was weighed and packed slowly in a column (2 cm I.D.) which already contained n-hexane. During the packing process, the column was checked constantly to insure that no air bubbles were trapped. After packing was completed, the column was eluted several times with n-hexane. Five grams of DCC were dissolved into 10 ml n-hexane and the solution was poured into the column, which was then washed with n-hexane a few times. A mixture of n-hexane and diethyl ether (Mallinckrodt A.R. grade) was then used to elute the column. The mole fraction of ether in the elut- ing solvent was increased gradually for each addition. Isomer A was eluted first and then the solvent was removed by a flash-evaporator. Isomer A was recrystallized twice from ether and then the crystals were taken out with ivory forceps to avoid contamination by metal ions. WE ar C: II me We SL1 dC an. She 47 Then the crystals were pumped on in a container on a vacuum line overnight. The white crystals melted at 61-62°C. If the m.p. range was larger than 1% degrees, the separation was repeated again using a newly-packed column. Isomer B could be stripped off from the column by eluting with anhydrous methyl alcohol (Absolute, Mallinckrodt A. R. Grade), then recrystallized from n-hexane (m.p. 69-70°C). II. SOLVENT PURIFICATION AND SAMPLE PREPARATION. Propylene carbonate (Aldrich), acetone (Fisher), di- methylsulfoxide (Baker, Reagent grade) and acetonitrile were first dried with calcium hydride under reduced pres- sure and then transferred by vacuum distillation to newly activated molecular seives (Linde Type-4A,(Matheson Coleman and Bell)) through a T-shaped joint with a coarse frit as shown below. % (15mm Fisher Porter joint and Ire/Penton coopllng 1's fid:".- I I. . ’ Ca H2 molecular selves 48 After 12 hours this solvent was vacuum distilled to another container (Figure 8). Before vacuum distillation, this bottle was pumped on the vacuum line for a few minutes through joint (1) and heated gently to get rid of adsorbed moisture on the surface. The distilled solvent was kept in the drying vessel and transferred to the sample tube when needed. The sample compartment was first pumped on line through joint (2) for a few minutes before transfer. After the desired volume of solvent had been transferred, the solution was cooled with dry ice and the sample tube was flame sealed off. During transfer, the pressure in the stock bottle must be higher than that in the sample tube; otherwise "bumping" of the solution can occur. Pyridine (Fisher) was dried over CaH2 under reduced pressure overnight and then further dried by the following method: Sodium and potassium (J. T. Baker Co. 99.99% pure) were packed into small ampoules as described in F. Tehan's thesis (89). A three-to-one ratio of potassium and sodium in the small ampoules was put into the sidearm of the bottle and sealed as shown in Figure 9. Benzo- phenenone (Matheson Coleman and Bell) was placed into the vessel in an amount slightly less than the stoichiometric amount of metals, then the vessel was vacuum-line pumped. When the pressure in the line went to 10'6 torr, the metal was heated by a gas flame until all metals left the first section and glass ampoules: leaving some impurities in 49 Kontes valve (1) TO vacuum TO vacuum (2) TV ‘ \J CIEEN ’" 5 mm Fisher-Porter and Penton Coupling l' ‘l Figure 8. Apparatus for storing solvent in the preparation of NMR sample. thi PUr off Sue] 50 constriction 1:12;) I‘d/K v v w benzophenone Figure 9. Vessel for Solvent Purification. this section. The first constriction to isolate the im- purities away from the Na/K alloy was made by flame seal- off. The sidearm was repeatedly heated and sealed off, in such a way the metal was finally distilled into the bottle. After the Na/K alloy was in the bottle, the last constric- tion was sealed-off and pumping was continued for awhile and then pyridine was distilled over. When pyridine was distilled through the vacuum line, a blue-colored solution (caused by the formation of benzophenone radical anions) indicated that the solvent was dried. The above method had to be modified for drying DMF, since it decomposes when pl TH pc n .1 po si reln 51 placed in contact with an alkali metal in basic solution. An anthracenide radical anion was first formed by distilling THF into the bottle which contained freshly distilled potassium and anthracene. A blue solution formed indicat- ing that the anthracenide radical anion was present. After pouring the blue solution into vessel B (Figure 10) the side bottle with metal was sealed-off. Then the THF was An Figure 10. Vessel for DMF Purification. .removed by vacuum distillation and the blue salt was Pumped overnight to make sure that all THF was removed. I‘~7,N-dimethylformide (DMF) (Burdick and Jackson Laboratories, Ir ar ex bl di wi Ph PP II in am loe 60 of Ste Us The mix is sin ”or: etel 35, 52 Inc.) was dried over CaH2 under reduced pressure overnight and the dried DMF was distilled into the vessel after evacuation to 10"6 torr. If the solution was greenish- blue, it indicated that the solvent was dry; if not, the distillation was repeated. All solvents except acetone were analyzed for water with an automatic Karl Fisher Titrator (Aquatest II) from Photovolt Corp. The water content was always below 100 ppm 0 III. THE NMR SPECTROMETER; MEASUREMENT AND DATA HANDLING. A Varian DA-60 NMR spectrometer was modified to operate in the pulsed mode at a field of 1.409 T. The 133Cs reson- ance frequency at this field is 7.8709 MHz. The field is locked by a home—built (90) lock probe which uses the DA- 60 console to lock on the proton resonance. A Block diagram of the spectrometer is shown in Figure 11. The spectrom- eter consists of three main parts. The first is a tuned transmitter/receiver section which operates at 56.44 MHz. The second section consists of a network of double balanced mixers coupled to a frequency synthesizer. The third part is a wideband transmitter/receiver network coupled to a single coil tunable probe. By utilizing the mixing net- work, broad band amplifiers and tunable probe, this spectrom- eter can observe NMR signals in the range of about 2 to 35 MHz. 1 53 .uoumfiouuoomm monocomou owumcmsE Hmoaoccfiuase on» no Esumsac xooam .HH masses 1------'--"-l-"-‘U----ll"U'-|'-"'---‘-|-'|-"--'I'-‘. I' I I I mw>_w0mz mmooa 55 E mgr. V7 momfizse ” . . . e A _ 553525. . 5528:“. 55:. . mmsa 26.. .r one": 02:22 a - .m 33.55. 36 OER A24 ,Ill... 20a“. 00 13a . > 3% as. s: o had—am Edam £62- > can 5&3 zmmaom u--------;- _ 4¢O¢ d2< V ¢w>_w0w¢ OwZDh N12 Qn l-----!...-.:...-h.--- KOhOmhwo wmdxa :90 2m hm>m noo— § 5300;. ”5.53),: OmZDh — .Zm —_ J a dc Oh m0m «.22 ¢nm< o>wusmmc may as omswmmv me guess .uco>H0m mo mmfiuuomoum owaflcmooaosc ecu mmoumxo o» kuoesusm HsowuHmEo eds omm.on smm.o H.4H m.pm Azomzv maeuuecoumos oma.on ems.o H.mH o.ms load musconuso ocoammoum msm.oa oss.o c.5H e.o~ weapons mom.ou mem.o s.s~ He.sm Anzac osssssuomasauosan Hs~.ou mos.o m.m~ ms.ss Romany osexomasmasaumEAQ s-.ou «as.o H.mm os.~a masseuse lemme osuso sueaeneuaoomsm nonesz uocoa ussumsoo uco>aom no GOHHOOHHOU UHHfimfin—HOAV m . £60550 Owufiomdwwn M so auwHabHumoomsm owumsmsz you coauoouuou can mucm>aom mo mofluummoum .osusn assesses .5 sense 56 paper tape as octal numbers. This paper tape was then read and punched onto computer cards in octal form. The CDC 6500 computer program CONVERT transferred the octal numbers to decimal numbers in a format which is compatible with the KINFIT program. The program CONVERT and its deck structure are listed in Appendix c. The equilibrium and kinetics data were all fitted with the appropriate equations by using the least-squares curve fitting program, KINFIT. The related equations and deck structure are listed in the Appendices. CHAPTER III STUDY OF FORMATION CONSTANTS OF CESIUM TETRAPHENYLBORATE ION PAIR AND OF COMPLEXES WITH MACROCYCLIC LIGANDS IN VARIOUS SOLVENTS 57 58 I. INTRODUCTION Previous studies in our laboratories (92-95) and elsewhere (75,83,96,97) have shown that alkali NMR offers a very sensitive probe of the environment of ions in electrolyte solutions; for example, sodium ions in various solvents, and solvent mixtures. The linear relationship between sodium-23 infinite dilution chemical shifts and solvent Gutmann donor numbers illustrates the capability of this technique (98) for predicting donor abilities of various solvents. Of all the alkali metal ions, cesium- 133 has the widest range of chemical shifts and its shift is therefore most sensitive to variations in the immediate environment. In 1975 Cahen gt 31. (93) showed that nuclear magnetic resonance of the lithium nucleus could be used to measure the stability constants studies of lithium complexes in solution. The purpose of the study described in this chapter was to extend the NMR investigation of alkali salt solutions to cesium salts and complexes in solutions. It should be noted that in solvents of low dielectric constant and/or in concentrated solutions, cesium salts may form ion pairs or even higher ionic aggregates. Naturally, it was important for us to determine the extent of ionic association of cesium salts in nonaqueous solvents prior to the study of the complexation reaction. 59 II. INVESTIGATION OF CESIUM SALTS IN NONAQUEOUS SOLVENTS. The concentration dependence of 133Cs chemical shifts as a function of the salt concentration was studied for CsTPB,CsI and CsSCN in pyridine, acetone, MeCN, PC, DMF, 3 and DMSO. The results are shown in Table 8 and Figure 12. It can be seen from Figure 12 that only in the case of solutions in pyridine, acetone and MeCN did we observe the variation of chemical shift with concentration,which is characteristic of contact ion pair formation. Of the solvents listed in Table 7, pyridine has the highest donor number (DN=33.1) and the lowest dielectric constant (e = 12.40), yet the curvature of the plot indicates that in this solution there is relatively strong contact ion pairing. Cesium triiodide is known to be a strong electrolyte (99) and the virtual absence of a concentration dependent chemical shift in pyridine solution indicates that the extent of ion pairing is very small in this case. Cesium thiocynate has a very low solubility in pyridine, so that chemical shifts were determined at only five points in the concentration range of 0.007 to 0.0005 M. The chemical shift becomes more diamagnetic as concentration is decreased. Although the change in chemical shift is small compared, for example, to that for CsTPB, the curvature indicates that both CsTPB and CsSCN form ion-pairs, since the curva- ture changes dramatically as the concentration decreases. It should be noted that it is the curvature of the chemical 60 Table 8. Cesium-133 Chemical Shifts of Cesium Salt Solution at 25°C Salt:CsTPB CsSCN CsI3 Solvent:Pyridine Pyridine Pyridine Conc. (M) ppm Conc. (M) ppm Conc. (M) ppm 0.015 37.81 0.007 -35.10 0.02 -32.93 0.0121 37.18 0.005 -34.94 0.019 -32.31 0.01 33.0 0.003 -34.32 0.013 -3l.69 33.15 0.001 -32.46 0.008 -30.76 0.009 31.29 0.005 -31.69 0.007 -30.45 0.008 29.12 0* -30.76 0.002 -28.90 29.13 0* -29.41 0.007 26.94 0.006 24.46 24.01 0.005 22.29 0.004 17.95 17.34 0.003 14.23 0.0025 11.74 0.002 8.03 0.001 1.27 0.0005 -8.25 0* -15.64 61 Table 8. Continued Salt:CsTPB Cs13 CsI3 Solvent:PC PC Acetone Conc. (M) Appm Conc. (M) Appm Conc. (M) Appm 0.02 36.63 0.02 34.43 0.05 20.73 0.01 35.89 0.01 34.43 0.03 21.81 0.008 35.89 0.008 34.74 0.01 22.90 0.006 35.52 0.006 34.74 0.008 23.21 0.005 35.27 0.004 35.05 0.006 23.52 0.003 35.27 0.002 35.05 0.004 23.83 0.001 35.02 0.001 35.21 0.002 23.83 0* 34.02 0* 34.45 0.001 24.30 0* 23.96 62 Table 8. Continued Salt:CsI3 CsI3 izigf Solvent:DMSO MeCN Conc. (M) Appm Conc. (M) Appm Conc. (M) Appm 0.3 -70.47 0.3 -4l.31 0.015 -20.88. 0.05 -69.23 0.1 -37.59 0.012 -22.74 0.03 -68.77 0.05 -35.42 0.009 -23.77 0.01 -68.46 0.03 -34.17 0.007 -25.44 0.008 -68.46 0.01 -34.01 0.005 -26.93 0.006 -68.15 0.008 -33.86 0.003 -27.59 0.004 -68.15 0.006 -33.24 0.002 -28.89 0.002 -68.15 0.004 -33.24 0.001 -30.38 0.001 -67.99 0.002 -33.24 0* -34.03 0* -67.96 0.001 -33.09 0* -32.61 0* Data NOTE: 63 Table 8. Continued Salt:CsI3 C313 C313 Solvent:DMF MeOH Formic Acid Conc. (M) Appm Conc. (M) Appm Conc. (M) Appm 0.30 -3.38 0.05 37.43 0.3 12.34 0.10 -0.12 0.03 39.61 0.1 20.87 0.06 -0.12 0.01 42.25 0.05 23.51 0.04 0.34 0.008 42.35 0.03 25.06 0.02 0.34 0.006 43.64 0.01 26.45 0.01 0.34 0.004 43.64 0.008 27.07 0.008 0.49 0.002 44.42 0.006 27.23 0.006 0.49 0.001 44.88 0.004 27.54 0.004 0.65 0* 45.03 0.002 27.85 0.002 0.80 0.001 27.85 0* 0.9 0* 28.23 * Data obtained by extrapolation. NOTE: CsTPB in acetone, DMF, DMSO are collected by L. L. Liu. Fieu 64 L. * — _----—g -—--m-==«_—.o;-_-.-:&:_-.o:_-:::-:.:t=.:.- "'7‘“ TPB 0 I5 TPB'__,° E e ————————— .---- o. __...o--“"" d "°__,_...-1:r"'" D" - c s - .. L1- ‘-----fi-—--——.—.&.--:__* ..... ---- '5 fi‘ '3 - scu‘ L. -5ot )- *TPB- "'1“ '3:' ‘7"- ‘."" '1'”"."‘7“"‘.“"T"‘.‘""I—T '3 . . . 1 * l..L.LlU 0.005 ‘ C ‘4] 0° PYRIDINE ------- MeCN ................. pc __ __ .__ DMSO ——--—---- ACETONE ._..__.__ DMF Figure 12. Concentration dependence of the Cs-133 chemical shifts of cesium salts in various solvents. shift that the ( shifi value show: CsSC) pyric and J the j The j in t} iOn I One giVeI ”hare the fa the 65 shift !g_concentration together with the limiting shift that determines the association constant. Thus even though the CsTPB curve shows a much larger change in chemical shift than does CsSCN, the latter reaches its limiting value at lower concentrations than does CsTPB. iThis clearly shows that the ion-pair association constant is larger for CsSCN. The infinite dilution chemical shifts of Cs+ in pyridine were -15.6, -30.8, and -29.4 ppm for Tea“, SCN‘, and I3, respectively. The value -29.4 ppm was chosen as the infinite dilution chemical shift of Cs+ in pyridine. The ion pairing constants of CsTPB and CsSCN were calculated in the following way. The exchange between free solvated cesium ion and the ion pair is fast on the NMR time scale, consequently, only one resonance signal is observed and the chemical shift is given by the expression (97) = x +X Obs fo ipdip = 1p where 6f and 61p are the chemical shifts characteristic of the free and ion paired cesium ion respectively, while Xf and xip are the corresponding relative mole fractions of the two cesium species. Obviously, whe sti The mat Sul ree to St. tt 91 T) 66 where CfM is the concentration of the free cesium ion and CtM is the total concentration of the cesium salt. Sub- stitution of Equation (111.2) into (111.1) gives M Cf 60138 = E—E (6f - 61p) + in (111.3) I: The concentration equilibrium constant for the ion pair for- mation is M M + - c - c 2 (111.4) Substituting Equation (111.4) into Equation (111.1) and rearranging gives {-1 + (1 + 4KainCtM)1/2] 1P 2 M 19 2Kayi C 6 obs t Equation (111.5) relates the observed chemical shifts to the total concentration of the salt (CtM), the chemical shift characteristic of the free Cs+ ion (6f) and that of the Cs+ ion in the ion pair (dip), the mean activity co- efficientyt and the ion pair formation constant Kip’ The values of CtM and 6f are obviously known and those of Y: were calculated using Debye-Huckel Equation (111.6) (3_= 5.3 A (100)). Equation (111.5) cannot be solved directly since it has two unknowns 6ip and K1 The values p. of dip and of Kip were obtained with the help of a non-linear lea The the be the Tab // 67 least squares curve-fitting program KINFIT (101). 6 1.823 x 10 3/2 |z+ z_|/I (DT) -log y: = (111.6) 9 1 + 5.029 x 10 a ,3 (DT)1/2 ‘— The results are shown in Table 9. The ion pairing constant (Kip) for the thiocynate salt is over twice as large as that of the tetraphenylborate salt. This phenomenon may be caused by the different extents of polarization of the Cs+ ion by the two anions. Table 9. Ion Pair Formation Constants of Cesium Salts in Various Nonaqueous Solvents. Solvent Salt Kip pyridine CsTPB (3.7:0.3) x 102 CsSCN (9.3:0.2) x 102 Acetone CsTPB (2.110.3) x 10 MeCN CsTPB (3.8:1. ) x 10 PC CsTPB (1.610.?) x 10 DMF CsTPB m0 DMSO CsTPB m0 III. have t Cs+ ic and 10 and co a POpu Shift a free Cs IESpect ComPle reaceic rEPreSE Ce. 68 III. FORMATION CONSTANTS OF CESIUM TETRAPHENYLBORATE COMPLEXES WITH CROWNS AND CRYPTANDS IN VARIOUS SOLVENTS. In a solution of a cesium salt and a ligand we may have the following solute species: ligand molecules, free + ions, ion—paired Cs+ ions, complexed Cs+ ions and, free Cs and ion-paired anions. When the exchange between the free and complexed Cs+ ion is fast on the NMR time scale, only a population averaged chemical shift is observed. 5 + 5. 5 x (111.7) obs = GFXF 1pxip + c c In the above equation 6obs is the observed chemical shift and Xf, X. 1p' free Cs+ ions, ion-paired Cs+ and complexed Cs+ ions Xc are the relative mole fractions of respectively. The corresponding chemical shifts of free, complexed and ion paired Cs+ ions are 6f, 6 6 The c' ip' reactions which take place in a CsTPB solution may be represented by: Cs+ + solvent I Cs+'solvent Cs+ + TPB- + Cs+'TPB- 4. Cs+ + ligand I Cs+-ligand. Cesium ions are not as strongly solvated as smaller ions actio that large compe be 51 CdUSe We Wr Where WQr‘ as EC. (21 69 ions and therefore, the ion pairing and complexation re- action may be more important. In this study it was found that the formation constants of the complexes were much larger than the ion-pairing constants: therefore, the competitive reactions of ion-pairing and complexation can be simplified by assuming that complexation is the major cause of the variation of the chemical shift. That, is we write _ v obs — GFXF + GCXC (III.8) where 0: >4 ll 6FxF + 61px ip The formation constants of the complexation reactions were determined by measuring the cesium chemical shifts as a function of ligand/Cs+ mole ratio, then followed by a computer fit of the data with the equation _ M _ L_ 2 L2 2 _ 2 aobs — [(K ct K 0t 1) + (K ct + K cfiz 2K c EC: L 1/2 Gf-ac + 2K ct + 2K cM + 1) II“ M] + 6c , (111.9) t 2K cM t Equation (111.9) has two adjustable parameters, the forma- tion constant K, and the limiting chemical shift of the complex 5c (for details see Appendix B). In the above 70 M t the metal ion and ligand respectively. Due to solubility equation C and c: are the analytical concentrations of limitations, the concentration of CsTPB was kept at 0.01 M for all complexation studies. The crowns used in this study were l8-Crown-6, dicyclohexyl-18C6 (DCC), and dibenzo- 18C6 (DBC). Their cavity diameters are reported to be in the range of 2.6-3.2 A (16). The cryptands used were C222 and C222B where B indicated the benzo group on the ther chain (Figure 13). The cavity diameters of the cryp- tands are about 2.8 A (102). All the results of this study are given in Tables 10-14 and shown in Figures 14-18. Cesium NMR resonance signals are usually very narrow (E 2H2). Therefore no error bars are indicated in the chemical shift XE! mole ratio graphs. The corresponding K value from each mole ratio study was computed by using KINFIT and is listed in Table 15. Two techniques were involved to compute K values with KINFIT: 112;! curvefitting and simulation. When the K 5 or larger, it is difficult to fit the value is about 10 region around the sharp break at the stoichiometric mole ratio. Usually the fitting process will either fail to converge or will give a very large standard deviation. The simulation technique uses KINFIT to imitate a system which is fully defined by a set of given conditions. For example, the initial estimate for a curvefitting process is the starting value which is varied by iteration. For A /O/-—-V—\O\ A /__\ \ O M 71 Amused “12$.an .. mum m nxuo . .mocsumwuolnmu was mc3onu .mH ousmwm A mummo V QZEUINNNIONmeOZOE _ r\ Anomo V ®IZ>>OmUImFIJ>XwIOJ O>O_Q leEOmUImFIONmeE n21»? R/lxwlxob , \ \nlJ,\(|/ o o ZQ O .66. “mosey slzsomolme 72 Table 10. Mole Ratio Study of 18C6 Complexes with Cesium Salts in Various Solvents by Cs-133 NMR at 25°C. Salt:CsTPB Solvents:Pyradine Acetone PC |18C6| |18C6| |18C6| [08+] Appm [08+] Appm [08+] Appm 0 32.38 0 35.41 0 36.16 0.28 17.66 0.36 23.77 0.27 28.80 0.51 9.37 0.72 10.42 0.5 21.44 0.78 -3.48 0.90 7.19 0.77 14.73 1.0 -9.91 1.00 6.35 1.00 10.72 1.19 —6.56 1.46 9.89 1.28 9.79 1.46 -0.222 1.19 7.47 1.58 10.54 1.90 5.65 1.80 13.15 1.74 10.54 2.04 8.91 2.00 14.46 2.12 11.93 2.56 15.90 2.50 18.28 2.60 13.42 3.03 20.08 2.77 19.67 3.10 14.54 4.09 26.06 5.15 28.99 5.40 19.29 6.28 34.06 7.08 33.18 6.89 21.53 6.91 35.45 10.3 36.81 8.90 24.51 12.62 40.95 16.54 40.54 22.40 33.27 18.95 41.38 31.3 35.88 73 Table 10. Continued. DMF MeCN DMSO [18C6] [18C6] [18C6] [Cs+] Appm [Cs+] Appm [Cs+] Appm 0 0.81 0 —24.51 0 -67.96 0.33 -0.37 0.33 -21.16 0.23 -58.18 0.63 -l.98 0.62 -18.55 0.59 -44.76 0.90 -2.73 0.77 -l7.15 0.90 -34.98 1.00 -2.82 1.00 -15.48 1.00 -33.40 1.35 -2.82 1.28 -14.45 1.24 -30.51 1.45 -2.63 1.40 ~14.18 1.45 -29.21 1.69 -2.54 1.89 ~12.96 1.80 -27.25 1.99 -2.17 2.06 -12.68 2.19 -25.95 2.72 -l.42 2.49 -ll.75 2.33 -25.67 3.0 -0.96 3.23 -10.26 3.12 -24.08 5.19 1.37 4.85 -6.81 4.72 ~22.22 7.60 3.70 7.26 -2.62 7.24 -20.08 10.23 5.94 11.05 2.78 9.85 -18.50 30.74 18.23 19.7 -12.81 74 Table 10. Continued. Salt:CsI CsI CsCl Solvent : DMF H20 H20 |18C6| |18C6| |18C6| [Cs+] Appm [08+] Appm [Cs+] Appm 0 -l.89 0 -0.55 0 -0.19 0.32 -3.01 0.28 -0.65 0.5 -0.19 0.51 -3.10 0.51 -0.65 1.0 -0.19 0.82 -4.03 0.72 -0.74 1.5 -0.59 1.00 -4.31 1.00 -0.74 2.0 -0.59 1.55 -4.03 1.48 -0.74 2.5 -0.75 1.81 -3.75 1.26 -0.93 4. -0.90 2.15 -3.47 1.82 -1.11 6. -1.06 2.67 -3.01 2.05 -l.11 3.15 -2.35 3.20 -l.38 8. -l.37 5.08 -0.40 2.49 -1.30 6.90 1.47 7.31 -l.86 4.06 5.19 11.41 -2.16 20.7 11.71 24.35 13.67 75 Table 11. Mole Ratio Study of BBC Complexes with CsTPB in Various Solvents by Cs-133 NMR at 25°C. Solvent: Pyridine Acetone PC L9§§l Appm LQEEl' Appm L2§§l Appm [Cs ] [Cs ] [C8 J 0 32.23 0 35.52 0 36.01 0.15 28.77 0.06 34.90 0.19 34.28 0.46 20.46 0.32 30.19 0.46 30.55 0.65 17.48 0.63 26.72 0.685 28.82 0.99 13.88 1.00 25.10 1.00 27.08 1.29 15.62 1.38 26.96 1.20 26.71 1.50 17.97 1.49 27.58 1.39 26.71 1.92 24.92 1.79 29.69 1.40 26.71 1.98 25.30 2.00 30.81 1.55 27.20 2.18 33.49 *2.91 32.30 2.17 28.20 2.41 29.02 2.24 28.32 3.22 34.11 *3.0 29.69 6.17 39.57 7.48 40.31 *9.81 40.68 *DBC not completely dissolved. 76 Table 11. Continued. Solvent: DMF MeCN DMSO L2§§l. Appm L2§§l. Appm LEEEl Appm [Cs J [Cs 1 [Cs ] ‘ 0 1.89 0 -24.41 0 -68.08 0.13 2.51 0.29 -20.07 0.13 -66.60 0.36 3.63 0.45 -16.84 0.32 -63.99 0.57 4.74 0.61 -15.23 0.55 -61.14 1.00 5.86 1.00 -8.90 1.00 -55.93 1.13 6.73 1.25 -5.18 1.14 -54.81 1.33 7.10 1.48 -2.57 1.20 ~54.31 1.54 7.71 1.59 -1.33 1.69 -49.60 1.86 8.59 2.04 3.13 1.84 ~47.74 2.06 8.96 2.27 5.62 2.14 —56.00 2.12 9.33 2.65 8.59 2.31 -44.76 2.61 10.82 2.95 10.95 3.33 11.82 4.07 17.15 3.74 12.44 5.82 23.36 77 Table 12. Mole Ratio Study of DCC Complexes with CsTPB in Various Solvents by Cs-133 NMR at 25°C. Solvent: Acetone PC Pyridine fiE§§% Appm £E§§% Appm %%§%% Appm 0 36.26 0 35.77 0 32.25 0.27 24.48 0.22 23.23 0.19 23.19 0.56 10.59 0.40 17.28 0.39 5.07 0.89 -6.29 0.75 —3.32 0.71 -l3.41 1.00 -9.14 1.00 -13.00 1.00 -30.78 1.15 -l4.97 1.10 —18.58 1.23 -33.76 1.44 -l8.45 1.50 -21.06 1.52 -34.13 1.87 -18.82 1.70 -21.19 1.64 -34.13 1.99 -18.82 2.14 -21.56 2.09 -33.76 2.17 -18.82 2.49 -21.68 2.13 -33.89 2.42 -18.57 2.97 —21.43 2.76 -33.02 2.83 -18.32 3.09 -21.43 2.93 -32.77 3.07 -18.08 3.25 -21.31 2.99 -32.40 5.86 -15.35 4.45 -20.56 4.53 -30.04 7.79 -13.61 6.29 -l9.32 6.11 -27.93 8.74 -17.84 9.35 —23.96 78 Table 12. Continued. Solvent: DMF MeCN DMSO £2991 Appm £2991 Appm [DCC] Appm [c.+] [es+1 [cs*] 0 1.15 0 ~23.88 0 -68.27 0.17 -3.07 0.22 -26.36 0.25 -65.17 0.46 -9.77 0.55 -30.58 0.45 -62.44 ‘0.73 -14.86 0.89 -34.67 0.72 -60.08 1.00 —20.32 1.00 -35.66 1.00 -57.60 1.25 -23.05 1.14 -36.41 1.26 -55.86 1.46 -24.66 1.65 -37.03 1.47 -55.24 1.60 -25.78 1.94 -37.03 1.89 -54.25 2.03 -27.39 2.02 -37.03 1.99 -53.51 2.19 -28.01 2.43 -36.66 2.6 -52.88 2.49 —28.63 2.75 -36.90 3.03 -51.77 2.91 -29.01 3.04 -36.78 3.37 -51.77 3.27 -29.25 3.71 -36.28 6.54 -49.78 5.06 -29.38 5.53 -35.29 8.52 -49.03 6.12 -29.50 6.74 -34.92 9.81 -29.38 8.86 -33.68 79 Table 13. Mole Ratio Study of C2228 Complexes with CsTPB in Various Solvents by Cs-133 NMR At 25°C. Solvent: PC Pyridine Acetone L—lfiif'i L—lizfi: L433? 0 36.34 0 32.25 0 34.90 0.21 22.74 0.32 15.37 0.36 14.38 0.52 -1.29 0.6 -3.61 0.5 -24.95 0.69 —14.71 0.83 -49.40 0.75 -39.47 1.04 -31.94 1.00 -66.89 1.00 -44.31 1.05 —35.11 1.16 -70.33 1.50 -50.14 1.16 -37.90 1.27 -74.83 1.70 -51.01 1.32 -39.67 1.36 -74.33 2.00 -51.75 1.49 -42.84 1.41 -75.33 2.50 -52.25 1.67 -44.87 1.64 77.56 3.53 -53.12 1.74 -46.06 1.77 -78.06 *w -54.0920.23 1.75 -45.73 1.99 -78.81 1.92 -46.75 2.52 -79.55 2.2 -48.05 3.20 -79.67 2.7 -49.45 *w -80.60:0.15 3.34 -50.2 *w -52.55:0.14 * Obtained by the curve-fitting program. 80 Table 13. Continued. Solvent: DMF MeCN DMSO |C22ZB| Appm ICZZZBI Appm [C2228] [03+] [cs+3 [03+] Appm 0 1.17 0 -24.09 0 -68.31 0.28 -4.12 0.35 -40.07 0.21 -68.46 0.41 -5.30 0.54 -50.62 0.53 '63.46 0.82 -11.20 0.68 -54.19 0.83 -68.46 1.00 -12.51 1.00 -65.51 1.00 '68.77 1.39 -17.07 1.32 -70.63 1.23 '68.46 1.53 -18.19 1.61 -72.33 1.42 -68.93 1.67 -19.77 1.98 -73.26 1.76 -68.93 2.0 -22.01 2.04 -73.26 2.18 '68.46 2.47 -24.99 2.41 -73.58 2.46 -68.46 2.90 '27.60 3.13 -73.89 3.07 -68.46 *m -51.8211.61 *W -74.-4i0.04 81 Table 14. Mole Ratio Study of C222 Complexes with CsTPB in Various Solvents by Cs-133 NMR at 25°C. Solvent: PC Acetone MeCN [c222] Appm [c222] Appm [c222] Appm [Cs ] [Cs ] [Cs 1 O 36.45 0 35.83 0 -24.40 0.13 17.37 0.30 -17.17 0.26 -97.14 0.48 -56.72 0.58 -102.26 0.51 -132.97 0.72 —113.68 0.89 -165.60 0.72 -166.31 1.00 -175.56 1.00 -182.52 1.00 -206.17 1.19 -184.87 1.18 -192.51 1.37 -208.81 1.45 -189.21 1.58 -199.42 1.55 -209.58 1.93 -191.38 1.86 -200.36 1.72 -209.74 2.08 -191.69 1.97 -200.92 2.27 -210.20 2.23 -191.54 2.42 -201.29 2.36 -209.89 2.93 -l92.62 2.89 -201.85 2.87 -210.05 3.17 -192.62 3.03 -201.85 3.11 —210.20 *m -193.75:0.16 *m -202.96¢9.19 *m -210.39:0.11 82 Table 14. Continued. Solvent: DMF Pyridine L—lizzi. —-‘E::i% ——‘:::i§ 0 0.80 0 -67.01 0 30.81 0.19 -13.31 0.25 -70.56 0.25 ----- 0.45 -35.02 0.57 -74.67 0.45 ----- 0.60 -41.85 0.92 -79.90 0.72 ----- 1.04 -63.30 1.00 -81.21 0.96 -217.33 1.20 -78.60 1.12 -82.52 1.28 -223.85 1.40 -87.60 1.46 -86.07 1.46 -224.47 1.70 -95.51 1.82 -89.99 1.81 -224.18 2.13 -105.90 2.12 -92.61 1.97 -224.18 2.40 -111.95 2.57 -95.97 3.05 -224.25 3.18 -121.25 2.75 -96.72 3.56 -124.51 3.26 -100.64 *w -155.80:2.30 *w -144.3413.52 1 Data collected by L. L. Liu. 83 ._ j 1..Cssph4 1n pv; II--C58Ph4 in MezCO; Ill--CsBPha in pc; .3 '1' IV--CSBPh4 in DMF; V--CSI in DMF; Vl--C58Pha in New; __€5() _y v11 --c:1 in H20; VIII-~C58Pha 1n onso __ 1D" 1 1 1 1 1 _1 L L 1 1 1 2 4 6 8 l0 4. 18C6/[Cs] Figure 14. Chemical shifts of cesium—133 as a function of mole ratio of [18C6]/[Cs+TPB’] in various sol- vents. 84 5C) ,_,..__...—i——-—o—-*—-u PYRI DINE /./ /. ,....~-y6*/ ACETONE PC APPM DMSO 1 1 l l I l L 1 2 3 4 5 6 7 8 9 [DBC]/(Cs+) Figure 15. Chemical shifts of cesium-133 as a function of mole ratio of [DBC]/[Cs+TPB'] in various sol- vents. [CsTPB]T = 0.01 M. 85 0‘. \‘K K‘u-mo—o-M"""-° \‘M-u-Hh -------------- APPM ,; A— f V 1 2 3 4 5 6 7 8 9 10 Figure 16. Chemical shifts of cesium-133 as a function of mole ratio of [DCC]/[Cs+TPB'] in various sol- vents. [CsTPB]T = 0.01M. 86 40 20 0' F" . . 0.. E 20, . . . &’ ~ InwF d -40~ ° _ . PC . ACETONE -60 1- 1;? DMSO 94 A". i A MeCN '80 ' V v PYRIDINE -100- l l l l l 1 2 3 4 5 lczzzsl [05" I Figure 17. Chemical shifts of cesium-133 as a function of mole ratio of C2228]/[Cs+TPB'] in various solvents. [CsTPB T = 0.01M. 87 4C 0 2 4 O. 0. Id \3. “100 ‘K\\* DMSO *4 cm '\ 04%, PC ‘200 °*D‘°*wo~~ ACETONE ""°‘"" MeCN " “1" W“ PYRIDINE l L L 1 l J l 031- N m 2 3 4 5 [cad/[05+] Figure 18. Chemical shifts of cesium-133 as a function of mole ratio of 5C2228]/[Cs+TPB'] in various sol- vents. [CsTPB T = 0.01M. .1 88 .vosuou xoaaeoo and on» mace «a an vouuasuauu ma co>am oaaa> on» can woman 03» once sawucuuauu aweuon you on n0>unu on» «0 nodes: on» .uo>mzo= .nuco>H0n anon» :« one sud: EuOu moxoamsoo Hun can and anon unnu casunoum aw an .nv .Nx onwauuuoo ou oHnaaao>u and snap susoco no: use ou>uoano aw newuaonmEou an. moaxx«.vo.n. ~osx.aflv. .61 n.ow..v a Na Anc~Amn a '2A was A a zoo: A . ..o«a . an o «AFN oax H.cflo.~ MA- N Any modxfia.ouu.acu ax owzo mo.o«vv.~u~x nflom «sax.fi.o6m.a. noaxflm.oflv.~. Anvmfion a noaxxnflm.u a use an. an. m.vo.vn u N: moaxas.owm.ns nadxxo.oflm.o~. H 4SA mSA sea A a acouoo¢ .61 “as «Am a «x «unconuou $2363.63 modified _ «afle nose oax.u.o“m.acu x ocoaaaoum an. “oax1~.ofln.~e HAH» a «x moHAAm.oAs.m. moaA H moaA nadxa~«.hv sea A x ocwcfiuam m-~u -~u .ouzuxsz. one coma ova .au.n~. auco>dom uaosuo> as nocnqu ruaz a: Ho.oH guano uo aucuuncou cosunsuou .mH canoe 89 simulation the initial estimate may be used to mimic the behavior of the system. For this reason, whenever the K value was too high to fit an experimental curve, the simula- tion technique was used to get a rough estimate of the association constant. However, such estimate cannot be taken very seriously, (i) COMPLEXATION REACTIONS WITH CROWNS The variation of the Cs—133 chemical shift as a func- tion of the 18C6/Cs+ mole ratio in different solvents is shown in Figure 14. It is immediately obvious that the solvent plays an extremely important role in the com- plexation process. The behavior in pyridine, acetone, and propylene carbonate solutions is especially interesting in that the Cs-133 resonance shifts linearly downfield until a 1:1 ligand/Cs+ mole ratio is reached and then shifts upfield as the concentration of ligand is further increased. The data seem to indicate a two-step reaction. First the formation of a stable 1:1 complex and then the addition of a second molecule of the ligand to form a "sandwich" complex (13) . In DMSO and MeCN solutions the Cs-133 resonance shifts only upfield with some indication of a weak "break" at the 1:1 mole ratio. The resonance continues to shift upfield even after the formation of a 1:1 complex but 90 there is no clear-cut evidence of a 2:1 complex. However, least-squares fitting of the data (Figure 19,20) based on the formation of only the one-to-one complex gave a large standard deviation and the fit was poor especially at lower concentrations of ligand. Therefore, a model was used which assumed that both one-to—one and two-to-one complexes are formed. The fit was much better as shown in Figure 19- Cesium tetraphenylborate complexes with 18C6 in MeCN showed similar features to those in DMSO. Once again, attempts to fit the data, as shown in Figure 20 by assum- ing formation of only the 1:1 complex gives large devia- tions of the calculated curve from the experimental one. An attempt was made to include the 2:1 complex in the calculation. This resulted in a reasonably good fit of the data but a big standard deviation for the first forma- tion constants (K1). This big standard deviation is mainly caused by the large value of K1. When 18C6 forms a complex with CsTPB in pyridine solu- tion, the change in-direction at 1:1 mole ratio clearly indicates the formation of a second complex. However, the variation of the chemical shift with mole ratio does not clearly indicate the stoichiometry of the subsequent complexation. In order to discriminate a sandwich complex (2:1) from a club sandwich complex (3:2), a study of the temperature dependence of the chemical shift was carried out. From the data listed in Table 16 and plotted in g s S S -‘- ' ' ‘ ‘I I o I! I "I I ' O °. (‘ c. I U ‘ o . u I ' I .. - C ' I I l O 4 I I 1 I I 4 '. 1 1 ' 1 1 o ' ' I '- o r ‘ ‘ 1’ l c ' 3 O 4 ‘ I . o ‘ , I I I o I : O ' n o ‘ ' I z I I U . ' U C a t ' I . I ' I .’ o 1': ' I 1 g ' ' I ' I ‘ C 111 1 ’ 4 5: £ '0...‘..0.‘.n0-.‘OO.-‘....g‘..Qos I..‘: -I‘ '.-~.0.0.’..OI'\oo-"unOoroOOO‘OOI0"...o'000‘oooo" “3.4-”. .1 1'1~ --- r1..41u-.10...cn|~.I4NI$( ' ° - t 0 (a) Cs+118C6 Model. :. OS. .05. ; II I ‘H , I 1 I t‘ o 0 I 4 I I 3 ' I 3 I I 8 S 1 o | : 1 ' : I ' 1 -' I f ! , l I I. I .I l .I ' ’14 l 1 I I n .' 1. I i .1 , ~ I O I 2 I I I 3 I o .. I i 1 ' I 1 1 a i 8 I .I I '. 5 i 0'. N '. ... I“ ... . ..'.....'..-...-.-. ..~.‘1.0--'--oo-.-.ooo'...n.‘oo.o~' Cd" 00..- 000000000000000 (\';~.0 ' .III ” :I' '7': I“ 1:“ ' J ;‘ 0’: 'I """"° Urn“! I '0‘..." b) Cs+ (18C6)2 Model. Figure l9.Non1inear curve fitting of chemical shifts vs. [18C6]/[Cs+TPB ] in DMSO. 92 m1mma+mo.._\_nmom.nn_ .wlm 852m 335050 «0 0503.1. 4.5.30 .H ~ . .Juooz mzo rp m2¢ .:Cu5..s .o..wg.r.x_ l O “:1 uqu—u—U‘ K II x C W K1 _ .1 :A m _ u a _ u m w a f C ~ ~ _ w a “c _x _ _w1111m1111f11110111111111U1111k1111611118111101111011110111-"11111-1nom-111:1111w . . . > cpblr >2 . <141: mrnv 0.. r. 2:. .r_ra orpqsrc4422. 7L... ...~..J........~U c. 4;; u: G .2L02 rrr. u bzwymnus_.mc.6_cm.1 u sc..oz mrp ha w243:._1.w1r.. u ~. u ~7rzuzu:_.::.x:r.. n htc_1 c:_ .2 y;4.2. .: n hum 00"” .zumz cw mmcfiacoz .ON musmam b.)." o~ kl —J0.HU ... ...n.00n.a¢0ano.a.u.uu"5006. 1......1 a '.'.l"." 0‘80'060'1'000I0000r0000 XX 8"IP 0'06l'--- 00-0~—0~‘0'~>~o-‘.o'0 u-no oo—ao cu HI—‘h‘b-‘I‘Iflnuhfi"‘h-O—o-‘h-‘u“HO‘O—OUOC‘ufih-b-«Itube—oh“.'DOboo—y—o.|-> .44»...w...._...un.u..w ..:.. vsnm.) u .ubrms_eunxu :« uzrmz a £25 w:» —d £241: .Duwwn 0C .—.fl.x W~.Jrv-1wb mtp .c w:.:: ....n:»... .4. .~._.. u. vm.wu 93 Table 16. Mole Ratio Study of 18C6 Complexes with CsTPB in Pyridine at 24°C, -6°C, -38°C. Appm (X§° 1:1 Complex) [1806] [03+] 24°C -6°C -38°C 1.00 0 0 o 1.34 6.36 13.80 20.01 1.45 10.24 22.49 31.95 1.77 15.51 32.41 45.78 2.00 18.15 35.52 51.83 2.20 21.71 40.01 .58.19 2.50 24.66 29.77 61.91 2.83 27.61 32.25 64.24 3.00 29.62 34.88 65.94 5.44 41.41 44.33 69.98 7.90 46.99 47.29 71.06 9.75 48.85 47.60 71.22 94 Figure 21, it is clear that the subsequent reaction forms a sandwich complex. The mole ratio plots in acetone, PC and DMF also exhibit a dip at a mole ratio of one. There- fore, by analogy, sandwich complex formation is postulated. However, one interesting feature in the mole ratio-chemical shift plots is different. In pyridine solutions the plot exhibits a sharp change in direction and a discontinuous slope. The other three solvents give a somewhat different shape. These differences reflect different values of K1 and K2. In pyridine, K1 is much larger than K2 but the latter is also large. Therefore, a sandwich complex starts to form only when most of the Cs+ ions have already formed the Cs+°18C6 complex. This condition gives a sharp discontinuity in the mole ratio-chemical shift plot at the 1:1 ligand/Cs+ mole ratio. For the other solvents, K1 is not as large as in pyridine and K2 is also smaller. Therefore, a "rounded-off" dip is produced, because at mole ratios around unity both complexation reactions (1.3;, formation of Cs+118C6 and Cs+- (18C6)2) are taking place. A typical fit of the data obtained in PC at 25°C with an equation which assumes a two step reaction with shifts in the opposite direction is shown in Figure 22. It is also important that, in spite of the great dif- ferences in the shape of the curves, the limiting chemical shift of Cs+-(18C6)2 complexes as determined by the KINFIT program (Table 17) seem to be independent of solvent. This contrasts with the large solvent dependence of the 95 .2 Ho.o n Hammeuuu nwmflowusm ca o.mm1 ..mu ..v~ um mma+mo \Hoowauv chum“ macs .65 puss» amussmso mmaussammo mo sauna 4 .HN mucosa mzaii z. mmvsmo<$oma O. m m h m m V m N _ a ~ fl a 4 a _ A a O 0. 1ON 10m V d d 1 w _ 10.? ~ . 6.6m .41 _. -om _ ~ .. 10m 601 ll |I11 . .- u 4. 111111111111 omml Li 96 .cowuowon moum103u n no mCfluuwm Hmowmmu ¢ .NN ounwfim p5u2_ uc or“ omU¢oumuoUQ cocooooaoauooouaoo ooooo ozu m----m1u--m---1m---1V1--umu-u-011--m1111w1111r1111w1111m1-11w11-1v»-11m1111r1111m1111m1-11m1--nm--.- I" u w ‘ v u - x u o . . . I _ x K o O A n .v u u u u u A u n m m W m m .u m m m n m .u m m m m. m m m u 1 > «bqwo >0 K (puma wz470. hZ~OQ OUF P.~oowwonn u no» m1» h< ura<> .o~mu¢ do .~.m.x m— J NaumtnN- n b.1mw1mu2~.oo.ucn~s N hana my; #4 w.._.—<>o nhkwq UIF h< UDJ<>oguom<>vSK—CO .WWWVMZM—wmvoanv1 :2 p m «~40 97 Table 17. Limiting Chemical Shifts in Various Solvents. Solvent Cs+ Cs+-(18C6) Cs+-(18C6)2 PC 36.45 8.1:0.2 46.5:0.9 Acetone 35.83 6.35 47. t9. Pyridine 32.38 -9.35 48.2:0.2 DMSO -67.96 -23.7¢0.4 49. :24. chemical shifts of the free cations and of the one-to-one complexes. This observation indicates that the primary solvation shell of the Cs+ ion is more or less completely displaced by the ligand when the two-to-one complex is formed. The attachment of two benzo groups on 18C6 to form DBC, was expected to result in weaker complexes. This aromatic group decreases the basicity of the oxygen atoms (10) and also decreases the 0-1-0 distance which causes a decrease in cavity size (50). complexes with CsTPB in six solvents was made. A mole-ratio study of DBC The results showed some similarity to the case of 18C6: i.e., in pyri- dine, propylene carbonate and acetone we also observed 2:1 complex formation, while in DMF there was no change of direction, but only a continuous shift to higher fields which continued up to the solubility limit of DBC. In 98 MeCN and DMSO solutions, although the shift is in the same direction as in the case of 18C6, the curvature of the plot is much smaller. This shows that the complexation is weaker. An investigation of the half-height linewidth (Avg) at room temperature shows that DBC also forms more labile complexes with Cs+ than does 18C6. In the exchange region, i.e., when the mole ratio of ligand to total Cs+ is approximately 0.5, the half-height linewidth for 18C6 in pyridine (36.59 Hz) is much larger than it is with DBC (400 Hz), while for C2223, A08 in this region is 15 approximately 15 Hz (Table 18). Another interesting fact is that in DMSO with C222B, there is no chemical shift variation as more ligand is added, while with C222, weak complexation is observed. On the other hand, all crown ligands form relatively strong complexes in DMSO. Perhaps this is because the solvent can still play an important role when crown complexes are formed. When Cs+ is com- plexed by cryptand but before it has been completely surrounded by the cage, a strong competition between ligand complexation and solvation may be taking place as illus- trated in Figure 23. Dimethylsulfoxide is a good donor solvent, so that a weaker complexing agent such as C2228 cannot compete for Cs+ and no chemical shift variation results as the mole ratio is increased. However, in the 101 'k Table 18. Half-height Linewidth (Av15 Hz) of Cs+-Ligand Complexes at 24°C. Ligand Ligand/Cs+ Avg (Hz) DCC z0.5 69.65 DBC zO.5 1 C222 20.5 >400 Hz C222B zO.5 14.47 18C6 40.5 36.59 0 1.00 0.39 16.17 1.00 2.22 1.5 51.00 2.0 51.00 2.5 51.00 * Av% = (v%)sample - (Vk)ref° ..' ' .- . ' 9 ‘ 5" . {.7 . o I . J'.‘ . . '. '4". ' .-'.'-' "'1‘ '-' 7:’.1:’.J"‘) 1% "I Q ‘ '3“. .0.- '\ 5.:0. ,_ int; W 5"?“ . . .. .3; : '0] ‘I 4' ' 3:0: 0" ft: 0:. 3.... . I..'..-. .0. ... , . u. . o :0 . a 41.. ° '3': no 0 4 Q; .z-II.J.... '0'. ”1 o {'4 .‘o 0. I: '. ° 115.9" - 3 21.34%: 0 1.‘ ‘58.. 102 IB-CROWN‘G The interaction of Solvent Molecules With Complexed Cesium Ion. Figure 23. 103 case of C222, a stronger complexing agent, the Cs+ can be complexed more readily so that there is some chemical shift variation with ligand concentration. Because of the deli- cate balance between solvation and complexation, it seems that cryptands can serve as good ligands for the investiga- tion of solvent effects and ligand effects on the complexa- tion reaction. The expression of Das (Equation 1.15) for the chemical shift in an electrolyte solution, indicates that the extent of the paramagnetic shift is determined by the overlap integrals of ion-solvent and ion-ion interactions. This idea, together with considerations of size and complexing ability can be used to explain the direction and magnitude of the shift which occurs upon complexation as well as solvent effects in electrolytic solutions. When a complexa- tion reaction occurs in an electrolyte solution, there is competition between the donating ability of the solvent molecules, the counterions, and the donating sites of the complexing agent. This competition will have two effects on the variation of the chemical shift with the concentra- tion of the complexing agent. For example, Figure 24 shows graphs of the chemical shift (in ppm) XE the mole ratio in DMF for all five ligands. This example clearly shows the two features referred to above. The "sharpness" of the curvature tells how easily the complexing agent can compete with the solvent, leading to displacement of the solvent 104 50 _ 93C 18C6 .__”_.u£p—~—""W—O 011" I III 4"”k 1 t " . g C2228 0. d -50 .— O 1 “100 " C222 L J L 1 J l L 1 1 I l 5 1O LIGAND/[cg] IN DMF Figure 24, Mole ratio-cesium-l33 Chemical Shift Study for Various Ligands in DMF. 105 shell and the counter ion. That is, the curvature gives us thermodynamic information. However, the direction of the chemical shift variation and the ultimate extent of the shift depends upon how much the donating electron pairs of the complexing agent can perturb the electron cloud of the cation. This perturbation, when greater than that of the solvent, leads to a paramagnetic shift. Dibenzo-18C6 appears to have weaker interaction of the lone pairs with the cesium ion than is the case for the other crowns, probably because of its rigidity and lower basicity of the oxygen atoms. In DMF solution, the chemical shift moves upfield with an increase in the mole ratio (that is, as the concentration of BBC increases). This phenomenon implies that the donating ability of the lone-pairs on oxygen in BBC is weaker than those of DMF. However it should be noted that repulsive interactions at short range contribute to the shift so that one cannot equate the magnitude of the shift only with an attractive interaction. Indeed, when the interaction is "forced" as when Cs+ is inside of a cavity which is too small, the repulsive inter- actions can lead to large paramagnetic shifts. In any event, the chemical shift variation indicates that complexa- tion is taking place. In the case of 18C6 which shows a change in direction of the shift at the 1:1 mole ratio, the upfield shift shows that the net contribution of the lone pair electrons of the 12 binding sites in a 2:1 complex is smaller than 106 that of the 6 binding sites of the 1:1 complex. This ob- servation probably results from a large decrease in the in- dividual overlap integrals in the 2:]. complex as the short- range repulsions are relaxed. The half height linewidth (Table 18) also decreases when the sandwich complex is fornmed. This phenomenon would be expected if the 12 oxygen atcuns were symmetrically distributed around the cesium ion. For DCC, the aliphatic cyclic substituent results in a better donating environment (_i_.£._, closer oxygen-Cs+ contacts) as manifested by the chemical shift-mole ratio plot. Among all three crowns, DCC has the largest para- magnetic shift even though it does not form the strongest (unmplex. The second formation constant of DCC complexes ‘With Cs+ is the smallest of all three crowns in a given SOlyent. Both cryptands, with two more binding sites, and Probably much closer oxygen-Cs+ contacts because of steric lfiJnitations can perturb the Cs+ electron cloud more and, therefore, introduce a larger paramagnetic shift than any (If the crown ligands. However, probably because of the rigidity and size of the tridimensional ligands the stoich- ixmmetry of the complexes stops at 1:1. A close examination of the K values for a given ligand shows clearly that the solvent plays an important role in the complexation. With crown type ligands, probably because the complexed (in 1:1 complexes) and the free cation both interact with the solvent appreciably, the 107 complexation constant is not as sensitive to solvent as with the cryptands. Figure 23 shows schematically how complexation and solvation might take place on opposite sides of the cation which could make this complexation system more affected by the solvent. Therefore, C222 was chosen to examine the relationship between solvent properties and the complexation equilibrium. The values of log K for C222 obtained in six solvents and in H O 2 and MeOH from another source (50) are listed in Table 19. Pyridine has the highest donor number but the lowest di- electric constant of all the solvents listed and it forms the strongest complexes. In PC, a solvent with the highest dielectric constant and the next to lowest donor number, an intermediate value log K is obtained. When water, DMF and DMSO are all solvents of high donor ability and high dielectric constant and all have small values of log K. These phenomena follow exactly the behavior found for ion pair formation. This fact may indicate that for a weakly solvated cation such as Cs+, the interactions between cations, anions, ligand molecules and solvent mole- cules depend on both the donor ability and the bulk di- electric constant of the solvent. The size and geometry of the solvent molecules may, of course, also play a role. l()8 ozu .uo>m3oz .ooEHOu xoadeoo and on» aaco aw mu poucasoamo ma co>am osau> on» can woman 03» ova“ acaumummou quuoo uoc on mo>uso on» we noomnm .muco>H0m omonu cw one saw: Shaw momedEoo Hu~ can and nuon ans» unannoua ma uH+ .mosac> oumewumm C mm.o III 4 I IIIII IIIII 4 N lows n o Inflow o o z .omo~.m lomov.q loqomo.m 15H.MM.M MM“ No.4 moo: Anal Away am. am. sv.oA any loo. o~.4 mm.m nm.o oA.AH1 A+Vem.a mA 1H1 zoo: ooo.o1 A~o .<.z mv.H o~.~ A+vvm.a vo.m AH. 0mzo on.o “my o~.H oa.~ mm.m A+om¢.~ om.m .Ho are #2 INS NeH. A~. mm.~ Amo vm.m no.4 vA Adv ma Adv mA AHV occuood Ha.1~. an. l~o Ho.o .N. o.ocontou 6H.m oo.m 4a 141 me AH. mH.q 1H1 ocoflmoooo meH.l~. mm.~ INC mmm.~ ANS ob.m o.4A m AAA. mm.n .H. o A 1H1 monsooso m-~o -~o lousuxfizv omo ooma oon .Auommv muco>~om msOwuo> cw mocmqu saw: madam Edwmou wo monao> x was .oH manna 109 IV. TEMPERATURE DEPENDENCE OF THE CHEMICAL SHIFT AND ITS RELATIONSHIP TO THE THERMODYNAMICS OF COMPLEXATION REACTIONS. (A) CESIUM TETRAPHENYLBORATE COMPLEXES WITH 18C6 IN PYRIDINE. In this section, a detailed study of the two step re- action of the 18C6 complexed with Cs+ ion in pyridine will be considered. The possible equilibria for this system were first prOposed as follows: When the mole ratio of (18C6) to (Cs+) is zero,ion pair formation can occur: K. Cs+ + TPB- $9 Cs+°TPB— (111.3) The observed chemical shift can be expressed as 6 X (III.4) 6 = 6 F F obs ipxip + When 18C6 is added to the cesium tetraphenylborate solution the possible reactions may be expressed as 110 + - ip + Cs + TPB -...:_ Cs -TPB- (III.3) + + 18C6 18C6 K! + — lp - + C8 ~18C6 'I' TPB =_——-.__=' TPB 0C8 ~18C6 (111.5) + + 18C6 18C6 + + - Cs 1(18C6)2 Cs -(18C6)2 + TPB In order to see if the above reaction scheme can be simplified, two studies of the concentration dependence of the chemical shift were carried out. The data obtained for reactions III.3 and III.5 are shown in Table 20 and Figure 25. It can be seen that the salt solution shows substantial ion-pairing (Kip =(3.7:0.3) x 102 as mentioned in Section 111.1) while the 1-to-l complex of 18C6 with Cs+ gives no evidence for contact ion pair formation. That is, K! = 0, and the above scheme can be reduced to three re- 1P actions, i.e. p Cs+:TPB- (III.3) +04 Cs + TPB- 111 Table 20. Concentration Dependence of the Chemical Shift at 25°C for CsTPB and its 1:1 Complex With 18C6 in Pyridine Cs¢4B Cs¢4B + 18C6 (1:1) Conc. 6conc. Conc. 6conc. 0.015 37.81 0.121 37.18 0.020 -7.25 0.010 33.15 0.015 -8.49 0.009 31.29 0.010 -9.58 0.008 29.12 0.008 -9.73 0.007 26.94 0.006 -10.35 0.006 24.46 0.004 -10.97 0.005 22.29 0.001 -10.66 0.004 17.95 0.003 14.23 0.0025 11.74 1.002 8.03 0.001 1.27 0.005 -8.25 112 .oswofluwm aw madam Edwmmo mo coaumuucmocoo map nufi3 coaumwum> umwnm a80fiEmno mMHIEdwmoo .mm whomwm :2. ZO_._. C3, [Cs+] is very low. For the third case, an arbitrary small number was chosen, [Cs+] z 10-7, as an to C initial estimate. Starting with an initial estimate, the first value of [C] was obtained and used to calculate a more accurate estimate of [Cs+]. This procedure was re- peated until the ratio of the "current" value of [C] to the previous value of [C] was close to unity. Then the final values of [C] and [Cs+] were used to calculate rela- tive mole fractions and thus Gobs‘ The Fortran expression and EQN subroutine deck structure which were used are listed in Appendix B-II. As the data listed in Table 1 indicate, most macro- cyclic complexation reactions are exothermic. It was 116 therefore of interest to investigate the temperature de- pendence of the complexation of Cs+ by 18C6 in pyridine. In previous work, Lehn and coworkers (41) and Cahen 93 31. (93) both noted that the apparent chemical shift of the complexed metal ion varied with temperature. However no detailed study was made,probably because of difficulty in finding a prOper absolute reference. For the present temperature study, an absolute reference was designed which was suitable for all temperatures (Figure 26). / TO VACUUM 5 MM A wanm 1% NMR was f .flONfll\NHJIAD NMR TUBE 0.5 M Cst AQUEOUS SOLUTION ‘ Figure 26. Insulated NMR Reference Sample. 117 A 0.5 M CsBr aqueous solution was sealed into a 5 mm pre- cision Wilmad NMR tube which was concentrically sealed in a 10 mm precision Wilmad NMR tube. The space between the tubes was evacuated to 10"6 torr and then sealed. The validity of this reference was tested and the results are shown in Figure 27 and are given in Table 21. The chemical shift of a 0.5 M CsBr aqueous solution in a 10 mm NMR tube was recorded as a function of temperature from the readout of the spectrometer. It was measured at the same temperature as the probe. The chemical shift of the absolute reference in the insulated tube was also recorded from the readout of the spectrometer. The measure— ment was made under exactly the same conditions as with the 10 mm NMR reference tube, but the insulated reference was left in the probe only during the time of data collection (ELEL' less than one minute). In this way, the temperature of the solution in the insulated tube is at or near room temperature. The constancy of the chemical shift for the sample in the insulated tube demonstrates the validity of this technique in providing an absolute reference. Of course, the results also indicate that the interaction of the solute with the solvent changes with the temperature to give a temperature-dependent chemical shift. Sharp and coworkers (83,84) also observed this phenom- enon, but they compared their chemical shift to an external reference which was measured at the same temperature as sample. Their major interest was to measure the variation 118 .mnsu mocmnmmmu pmumasmcw on» NO unme .bm wusmflm 0e mm3h O ow.Hm aN.Hm nv.mv mm.hv vo.mv NH.ov am.~n o~.mN mm.mH ho.m mh.mHI «III cIIII ou.NH +mm~ mN.om am.om mH.mv v0.0v oo.~v mH.¢n NH.HM mm.m~ v~.mH oo.w mm.mHI cIII «IIII oo.mH mmw wn.on mo.on vm.mv hh.vv mv.ov mo.wn vw.m~ n~.m~ hm.va oh.v mm.hHI cIII «IIII ha.ha va au.mv HH.mv ma.vv mo.av hv.mm on.nn mw.m~ Hm.HN om.HH Nm.n no.wHI aIII cIIII ~m.a~ mmm Ho.mv ch.mv sm.mn m~.mm nm.m~ aw.v~ oa.md mm.mH mm.h ww.a Hm.NHI vo.H mm.oa mh.mm New ~N.nv mN.Nv Hm.mn mo.m~ mh.~n no.ma na.na hN.HH mh.v o~.o vv.OHI mv.v mm.HH om.m~ mow o~.a -.m oo.v w~.m nw.~ vn.~ o.~ nh.H nm.~ mm.a o.H m.o mw.o o ongku vo.mm oo.mn «n.0m .m.n_ on. 05d nfio ow.HI mn.mI wm.m mo.va mm.H~ mm.~m haw mh.m om.h ve.m o.m m.~ ~.N o.N hh.H MM.H oo.a no.0 om.o MN.o o m Emu xo _oUmHH .dan ocwowuxm ca momfi uo mucmmmum ecu ca maemo mo unacm Hmoasmno osu How mono musuaquEmeIowuamumHoz .NN manna 122 A ppm .; [IBCG] IO [C5¢4d Figure 28. A three-dimensional plot of the cesium-133 chemical shift !§_mole ratio and temperature (°C) for solutions of CsTPB and 18C6 in pyridine. The concentration of CsTPB was 0.01 M. 123 estimate of K1 through simulation of the data by the iteration technique previously described. All such simu- lated "data" were compared with experimental results and adjustments in K1 were made as needed. The values used for K2 were obtained by a least-square fit of the data obtained at mole ratios higher than one. Two representative results are shown in Figure 29. The corresponding K1 and K values 2 are listed in Table 23. Table 23. The Simulated Formation Constants (K ) and Least- squares Adjusted Constant (K ) for the Complexa- tion of Cesium Tetraphenylborate by 18C6 in Pyridine at Various Temperatures. Temperature °K K1 K2 297 5 x 108 71.3 285 5 x 108 116.6 272 5 x 108 209 255 109 414 244 3 x 109 605 235 6 x 109 1140 229 3 x 1011 3440 Since the K1 values are much larger than K2, at mole O + 0 ratio equal to one, most Cs can be con31dered to have been + . . . converted to Cs ~18C6. Therefore, it is p0831b1e to obtain K2 values by fitting the data beyond a mole ratio of 1:1 with 124 E Q 0. G --::::::::::-;::;:W 235'K I 1 L g L L 4 .1 l 1 ‘30 23456789l0 [mod/[05+] IN PYRIDINE Figure 29. A Comparison of simulated data (0) with experi- mental data (A») for a chemical shift - mole ratio study at 290°K and 235°K for solutions of CsTPB and 18C6 in pyridine. 125 the KINFIT program (Table 24) and thus obtain the enthalpy of the second complexation reaction. This procedure gives H2=-6.l6i0.09 Kcal/mole. The relevant thermodynamic param— eters are calculated by the following relationships: @— : 1%?- (111.17) d(1/T) AG = — RTan (111.18) as = Af-‘g—AE (111.19) and the resultant values are (AG2)298 = -2.8310.004 Kcal/mole, A82= -ll.l810.31 cal/(mole K°). Again, it is worthwhile to note that the limiting chemical shifts of Cs+°(18C6)2 are nearly temperature independent. This phenomenon may imply that the first solvation shell has been completely displaced by the ligand and also that there is no pro- nounced conformational change between 24° and -44°C. (B) CESIUM TETRAPHENYLBORATE COMPLEXES WITH C222 IN VARIOUS SOLVENTS The variation of chemical shift with temperature was also studied for the cryptand—Cs+ complexation reac- tion. The implications of these variations will be discus- sed later but in this section the overall formation con- stants of C222 complexes with CsTPB in DMF, PC and acetone 126 us.eumeos.emo m.os~.eeu n «ma meosxemox eoo.osomm.~1umm~l~oae H no.0Hm~.¢ u mmmxmsaev «Hoaxemom H.¢H~.mI umma mH.Hm NH.OH 0.5 moexxefie.ee mm~.e mmm mm.em mo.osev.m Noexxm.o“o.oc mmo.¢ «em mm.em ~.os o.m moexxm.ose.ac mma.m mmm ma.mq mo.o“vm.m moexAH“H.~e eem.m New a¢.mv mo.oese.v ms.eee mom.m mew m~.mv ~o.ose~.e efl.ee emm.m new Ewaasquv «mad «M HIAmHsumummEmavxmoH Axov musuoummfime .mousumuomema msowum> um mcwpflumm CH mamomav.+mu H.mUm +mUmH.+mU mo coauommm on» How mnwumfimumm oesmascosumns 0:» one xmedsoo “somev+mo on» no umegm Hmoesmno mceueEee .em wanna 127 will be considered. The relevant thermodynamic parameters of the complexation reaction can be calculated from the temperature and concentration dependence of the chemical shift, just as was done for complexation by crowns. For this study two concentrations of CsTPB were used, 0.02 M and 0.01 M. The external reference was 0.5 M CsBr aqueous solution in a vacuum-jacketed tube. All of the chemical shifts measured at various temperatures are listed in Tables 25-29 and shown in Figures 30-32. The formation constants and limiting chemical shifts were computed by the KINFIT program and the results are listed in Tables 30-32. The corresponding thermodynamic param- eters are listed in Table 33 and plots of fin K v§_l/T are shown in Figure 33. An interesting observation, which has consequence to be discussed in Chapter V is that the limiting chemical shift varies with temperature and as the temperature gets lower, this variation becomes smaller. In addition, the chemical shifts at low temperatures approach a value which is independent of the solvent used. These data indicate that the C222 complexation reaction with Cs+ is not validly described by a simple complexa- tion reaction, such as Cs+ + 0222 I Cs+C: c222 in which CSEZZCZZZ denotes a complex with a Cs+ ion in the center of the cavity. More discussion of this anomaly will be given in Chapter V. 128 oo.H~NI mm.mH~I om.momI Hm.omaI mo.mbHI «m.~mHI mv.omaI mm.maHI oo.¢ IIIII mm.hH~I om.wo~I om.omHI Hw.thI va.omHI H5.mNHI wm.moHI oo.m IIIII mm.hHNI om.momI mm.mmHI mo.thI mv.omHI mH.mNHI ma.moaI om.~ oo.HN~I mm.nH~I om.mo~I hm.mmHI mm.m5HI m¢.thI mm.mHHI va.vaI Ho.~ mm.o~NI mm.mH~I om.momI mo.mmHI ¢~.H5HI mw.m¢HI mm.oHHI mo.omI on.a oo.HNNI om.>H~I mv.momI mv.mmHI m>.HhHI Hm.H¢HI mm.moaI mm.~mI om.a mm.om~I om.>H~I mm.momI hm.omHI mm.mmHI Ho.mmHI mv.mmI mm.mhI om.H no.0NmI hm.mHNI nm.¢omI om.thI mm.vmau mm.mHHI mm.m>u m~.nmI oo.H IIIII IIIII IIIII hm.~mI mo.m5I mm.mmI wm.mMI mm.mHI mm.o IIIII IIIII IIIII ah.oml mo.m¢I «H.5NI mm.oHI ma.o Hm.o IIIII IIIII IIIII -.mI vm.m mm.oH no.ma mm.om o~.o mm.m~ mv.Hm ow.mm mv.~¢ mm.mm mm.mv mv.mv ma.me o oOVI oHNI omI 6am owv own 6mm omoa Uo musumummama oHuom 0H0: .om cw ammo mo mocmmwum map Ge A: No.ov mmemo mo umwsm amuseoso map you mama mnsumumafimquwumm 0H0: manna 129 Hm.thI Nb.mvHI v5.0HHI na.m hm.vaI mm.~mHI eH.m mw.thI mv.mvHI hm.moaI mm.~ mm.mmHI mm.~mHI mm.~ mh.onHI hw.ovHI ha.mmI m~.~ hm.¢mHI >5.HmHI m~.~ mm.mmHI mo.meI mh.va mo.m hm.¢mHI mm.HmHI mo.~ ¢~.vaI mH.~HHI mo.nnI mv.a m~.¢mHI Hm.HmHI mm.H hm.mmHI wm.¢NHI «m.mmI ma.a mm.~mHI ¢¢.thI mv.H wm.veHI om.ooHI om.va o.H Hm.thI oa.thI mH.H hm.HmI mm.HmI mv.h~I mh.o vn.oHHI mm.mHHI ~>.o mm.th mh.HmI mm.oHI mv.o IIIII mm.mI mw.o m~.b¢I ov.omI mm.ho mm.o IIIII «H.5N mH.o mm.ma mm.m~ mv.vm mmH.o Hm.mv N~.wv o om.mv mo.m¢ -.m¢ o com 6mm oov och oooa Uo wusumumafioa owuom Uo musumanEma oHumm 0H0: mac: .Um cw NNNU mo moswmmum 0:» a“ A: Ho.ov mnemu mo amenm HMUHEmnu 0:» “0m sumo musuouomfima I ofiumm 0H0: .mm manna 130 hv.am~I mm.mH~I mn.mo~I Hm.mo~I mm.HmHI hm.thI mm.~¢HI mm.HHHI mn.ooaI mm.q oo.ammI mm.mH~I mh.mo~I mm.vo~I Hm.mmHI mH.HnHI no.>maI mm.moaI mh.HmI m.m IIIII IIIII mv.mo~I no.vo~I no.5mHI vm.mmHI em.m~HI om.omI mm.nhI m.~ h~.mN~I N>.FH~I mv.homI ~m.~oNI mm.mmHI ¢¢.mmHI hm.HNHI mh.va vm.H>I em.H IIIII IIIII mm.oo~I no.Ho~I Hm.mmaI mn.mmHI mo.mHHI mo.omI Nv.me n.a IIIII IIIII m~.mo~I mm.ao~I vw.HmHI mm.mmHI nv.moaI mN.NhI mH.omI mv.H IIIII IIIII om.Ho~I ~m.omHI om.vnHI mm.¢vHI ~m.mmI Ho.~mI ov.HmI H~.H o~.m~mI o.¢HmI mH.vaI mo.mmHI mm.~mHI mw.amaI mm.omI mm.mmI ah.va o.H IIIII IIIII IIIII mm.bmHI no.5mHI Ab.oHHI oh.o>I wh.N¢I mh.mMI mh.o IIIII IIIII IIIII m~.mmI m~.va «m.pnl mm.va Hm.m~I ho.H~I m.o IIIII IIIII IIIII vm.nHI om.o¢I mm.~mI mo.>HI mm.mI mm.vI ma.o oa.mHI Ho.mHI mm.oHI om.mI HH.>I m~.m mm.a mm.m mm.w o thI onI oovI om.mmI oHNI omI new oomw ooem Uo musuoummEmB ofluom 0H0: .mzo cw NNNU mo mocommum may as A: ~o.oe mmemo mo umenm Hmoesmso map Mom mama muoumuwdsme I oeumm 0H0: .em manna 131 om.mo~I m>.~man mv.moHI mm.mmHI Hm.v~HI mm.m nm.¢omI m¢.omHI mm.mmHI w>.mmaI m~.H~HI ma.m mm.mo~I mm.>>HI v~.omHI Ho.m~HI mm.HHHI v.m hm.~o~I ev.mnHI mm.mmHI ~.v~HI m.moHI ma.~ ~m.ac~I ~m.wmau vm.¢vHI aa.mHHI Hm.mmI h.H Hm.meI m~.~mHI wm.mvHI mm.moHI ow.hmI v.a >.HmHI mo.~mHI hh.omaI mo.mmI Hm.mnI ~.H vw.meI hm.m¢HI m~.qNHI -.mwl «m.HnI wo.H IIIII mm.mmI om.mhI mm.HmI mm.HvI v.0 IIIII mh.o5I o.me em.HvI mo.mmI m¢.o IIIII m.omI mh.m~I m~.mHI Hm.mHI ma.o ¢.oHI ¢.¢I m.mI vm.o om.o o omMI oma oNI oma omm Uo musuouomsma oeuom mHoz .mza me ~an mo monomoum 0:» ca A: Ho.ov mmBmu mo Madam Hmowaonu map How oumo musuoumdeme I owuom 0H0: manna 132 em.mm~I me.om~I vo.m-I m~.m~mI me.oH~I mq.~I~I ov.amHI mm.mmHI we.HmHI ~o.v em.om~I mH.ommI mm.m~mI m~.mm~I mH.oH~I m¢.~o~I He.mmHI ~m.emHI me.emHI Ho.m m~.mm~I IIIIIIIIII mm.m-I me.mH~I mq.mo~I o¢.mmHI om.emHI m¢.omHI om.~ mm.mmmI me.om~I mo.mm~I m~.m-I em.mH~I mm.eo~I ee.mmHI me.mmHI m~.meHI mm.e em.mm~I IIIII IIIII m~.m-I vm.mH~I H~.Ho~I ma.mmHI Hm.vaI mo.oeHI oe.e em.om~I IIIII IIIII em.~m~I mm.mHNI m~.oo~I Hm.mmHI «m.~mHI Hm.meHI om.H mm.mm~I ee.m-I ew.o-I em.-~I -.mH~I eo.mmHI mo.mmHI m~.meHI em.mmHI om.H em.om~I ve.mm~I mH.vNNI m¢.eHNI mv.vaI ee.meHI 5H.H5HI HH.¢mHI ~m.m¢HI oo.H IIIIIIIIIIIIIII IIIII IIIII om.H~HI me.mHHI mm.eoHI ov.ooHI He.o IIIII IIIII IIIII IIIII IIIII ee.eeI oe.qu vm.mmI oe.omI me.o IIIII IIIIIIIIIIIIIII IIIII e~.eHI He.o~I mm.~HI so.HHI mm.o ma.e~ mm.m~ me.om Hm.~m om.mv mo.ev ~m.ev mm.m¢ va.me o ova oveI ommI oHNI oNI ovm 6mm owe ovm Uo musumnmmsms oeumm mac: .msoumod ca NNNu mo mocommum may as A: No.0v mmemo mo umesm Hmoesmso on» now mono ousumumdswe I oeumm 0H0: .mw menus 133 W( PCS wfl -50. x. E \ ‘ I ‘ I 'Q -100L- 5 I |. ' 105° ’ I 'I' O : ‘ 96 I g . O -150' ' ’ 72° I ' I | h I o I 46 I I , II » .‘****—*-* -----l- -2° 0 . 133:: 22:22:: :1: :33. ' 1 l L. L k [c222] 1 2 3 4 5 11 (5‘7 Cesium-133 Chemical Shift vs Mole Ratio of 0222 to 03+ in Propylene Carbonate at various Temperatures. Figure 30. 134 O I $1 DMF I r I -50 I- \ I r , . I. \\ .~ P ‘x' 3 1. . I- '1‘ o -100"' a 54° :1 o o E L a I 46 t II )- I " 'I'I ‘ I! . 1) 2" n -150 - ‘3‘ ‘ II I- II N . b- “‘ 1' .. _2° | I- ‘3‘ . 1} - 1 -21° -200 L- I": . II ’40 L- II Ig___ r. 'I' “‘----——-o ————— -o- -58° I .. Luau-x ------- 46- ----- *- “13° L 4 lie 1 1 : [czn 1 2 3 4 5 [Cs Figure 31. Cesium-133 chemical shift vs mole ratio of [C222]/ [Cs+TPB’] in DMF at various temperatures. Figure 32. 135 Cesium-133 chemical shift vs mole ratio of C222 to Cs+ in acetone at various temperatures. 136 50 0 in i 9 -50 I- ' f! (. .400. ’ » ;: 1 . 1 II -450' i r "I 3 \\ . _ 54° . \. 46° —2m- a“ séuiue—t—i—g 32: . I \k x-II-x-mII-«II- ----- -II- -2° \H+--o--¢.-----.- .21" I o—o-o-«x-ouo-«ue- 44° bs-Ih-a-«r-a-ut-u-«o- -64° -250T 1 l l l a». Figure 32. #_ C222 Cs 137 Table 30. Formation Constants and Limiting Chemical Shift for the Complexation of CsTPB by C222 in PC at Various Temperatures. Temperature (Appm)lim K [Cs+]= 0.02 M 378 -129 11 (1.7 1.1)x102 369 -138.710.7 (3.610.3)x102 345 -156.110.6 (1011)::103 319 -176.910.3 (4.010.9)x103 302 -191.1810.04 (6.710.4)x103 271 -208.80 ------------- 252 -216.35 [Cs+]= 0.01 M 373 -184 12.0 '(2.610.2)x102 343 -202.910.5 (8.610.4)x102 313 -220.510.2 (4.110.3)x103 298 -230.210.2 (10.il)x104 138 Table 31. Formation Constants and Limiting Chemical Shift for the Complexation of CsTPB by C222 in DMF at Various Temperatures. Tempeiature (Appm)lim K [03*]= 0.02 M 327 -134.410.4 44.310.5 319 -136 12 70 16. 297 -156.410.2 148.1 2. 271 —178.2410.07 512.1 6. 252 -193.810.2 (11.110.8)x102 233 -209. 11. (6.12)x103 215 -219.5 ---------- 200 -231. ---------- [03*]= 0.01 M 298 -155 11 156 6 288 -159 13 (2.6i0.3)x102 271 -172.010.s (6.910.3)x102 258 -185.51o.8 (9.210.9)x102 139 Table 32. Formation Constants and Limiting Chemical Shift for the Complexation of CsTPB (0.02 M) by C222 in Acetone at Various Temperatures. Tempeiature (Appm)lim K 327 -183.910.2 (1.910.2)x103 319 -189.310.8 (3.810.3)x103 302 -199.710.2 (1.811.1)x104 297 -202.710.9 (1.810.5)x104 271 -216.15 ————————————— 252 —223.28 ————————————— 229 -23o.02 ............. 209 -236.57 ————————————— 140 Table 33. Thermodynamic Parameters for the Complexation of CsTPB by C222 in Various Solvents AH AG an AS Solvent (Kcal/mole) (Kcal/mole) (298°K) (cal/mole°°K) DMF -7.510.3 -2.9410.01 4.9710.08 -1511 PC —lO.9iO.9 -5.50i0.01 9.310.2 -18i4 *-lO.810.l -S.4 10.001 9.1910.03 -l8.110.3 Acetone -15 12 -5.910.2 9.9 10.2 -31i7 *[Cs+] = 0.01 M. All other cases [Cs+] = 0.02 M. 141 ”ACETONE an DMF l _L _L .1 3.0 3.5 4 0 1,6 (K") Figure 33. A plot of an E l/T for the complexation re- actions of Cs+ with C222 in acetone, PC and DMF. CHAPTER IV A STUDY OF THE DYNAMICS OF CESIUM TETRAPHENYLBORATE COMPLEXES WITH CROWNS AND CRYPTANDS 142 INTRODUCTION The complexation studies described above showed that the stabilities of crown and cryptate complexes strongly de- pend on the nature of the solution and that the topology of the ligand dramatically affects the extent of the complexa- tion reaction in a given solvent. Therefore, NMR line- shape analysis at various temperatures was used to study the kinetics of the complexation reaction. The rates of cesium ion exchange in the presence of the macrocyclic ligands in PC, pyridine, acetone and DMF were calculated from the exact expression for a general two site exchange of uncoupled spins. The equations were fitted to the ex- perimentally observed lineshape with the aid of the generalized weighted non-linear least squares program KINFIT. It was suspected that the activation parameters obtained from line- shape analysis at various temperatures might help to further investigate how the topology of the ligand and the nature of the solvent affect the complexation process. II. DETERMINATION AND INTERPRETATION OF THE LINESHAPES The modified Bloch equations (103,104) which describe the motion of the X and Y components of magnetization in the rotating frame, are expressed as follows: d6 1 _1 A = _. ‘ dt +0IG iyfllMo +1 G A A A B B " TA (1v.1) GA 143 144 "d? + aBGB = ’iYH1M°B 4' TAlGA ' T131613 (IV°2) with GA = uA + ivA (1V.3) where v represents the absorption mode lineshape and u represents the dispersion mode lineshape. Then G(m) is a general expression for the lineshape which can predict the entire range of exchange from the extreme slow limit to the O 1 f . f << ; extreme ast limit At the ast exchange limit 1 753:5;7 and only one signal is observed, while at the slow exchange . l . . . limit, 1 >> 13;:5377 one can distinguish separate signals from the two sites. The exchange times, 1, at different temperatures supply all of the kinetics information. (A) MEASUREMENTS IN THE ABSENCE OF EXCHANGE In order to determine the linewidth of the two dis- tinguishable sites at various temperatures in the absence of exchange, separate lineshape analyses were made of the salt (site A) and the completely complexed Cs+ ion (site B) at various temperatures. The values of TZB were studied over the temperature range where the formation constant is larger than 103. Therefore, the populations PA and PB can be determined directly from the mole ratios of ligands and cesium salt added to the solution. All experimental 145 lineshapes in the absence of exchange were found to be Lorentzian. The values of m and T2 for the free and the complexed ion were determined by fitting a Lorentzian line to the observed signal. A typical example for a fit of 0.01 M CsTPB in PC is shown in Figure 34. The five adjustable parameters were the amplitude, K, the Larmor frequency, w, the linewidth parameter, T2, the height of the baseline, c, and the zero-order phase correction, 8. The first-order phase correction, 81, can be determined experimentally by following the procedure in the NMR Manual from the Nicolet Instrument Corp. (105). Therefore, the first-order phase correction is introduced as an externally determined quantity while all of the other five parameters were adjusted by KIN- FIT. The factors which can affect these parameters are described in detail in reference 95. The variations in linewidths with temperature for the free and complexed cesium ion, in the presence of 18C6, DCC(A), C222B and C222 in PC solution are listed in Tables 34 and 35 and the data are plotted in Figures 35 and 36. The natural linewidth of the cesium ion is very narrow (since the quadrupole moment, Q = 0.003 barns, is small) and the linewidth is predominantly determined by the inhomogeneity of the field. Therefore, only T; values are reported. *1 df'db(T*’1-(T )"1 2 s e lne Y 2’ ‘ zinh where TZinh is caused by inhomogeneous broadening The quantity T ( “1 + T2nat) and T2nat is the natural linewidth of the species under 146 .UomeI pm mumconnmo mamaxmoum cw mnemo z Ho.o mosses» Isoo sofiuSHOm m mom wmonmmcfla mmaImU mo mflmaamcm BHmzHM Havana» m .vm onsufim psur~ no ozw coazaun oestooo-ooshouoouo:noose 02w m m n u n m w u u u n m M n w m I P m m 1 .u nun "nun oooo coco xxx x . «ix 5:; noco an 8 nun" In UK Gun no x Oux uxnn nx xx 0 H XX" 00 . u x x00 "K K O O x K O o o u xo . x o O K o b x o .. O x x o O x x o . v o u x u v o o x x O I x — O . x o x .u 0 . x o K v x O . o x xx 0 . o _mI m m m w m m m w w m m m w m----mII--m----mIIIImIIIImIIII > «same >c x «some uxcm uz. z~ um: ez~0d owpagauqau 92¢ 4ocoow~o~. u QC» wL» h< wDJ<> .cumwa mo ._.N.K m— 4¢Uwhmw> oc.w~.0~.o u. F7m3waU2_.oO-w¢a—u n FICA“ U1» hd wD..<>oJOowo¢—u H upkUI— “I... h< UDJ<> o.u.0<>¥.x GO A~umvx W~ MVOWUQ 147 £939 own emo.o I assayee “msoz Amoo.ov veoo.o m.mo~ Amoco.ov «Hoo.o o.mm~ Aaoo.ov moo.o m.qmm “moo.ov voo.o o.¢m~ Avoo.ov omo.o o.mv~ “Hoo.ov moo.o o.~m~ “voo.ov mvo.o o.mm~ Amoo.ov omo.o o.om~ Amoo.ov «no.0 o.vmm “moo.ov -o.o o.mm~ Amoo.ov who.o 0.5mm Aeoo.ov nmo.o o.mom Awoo.ov mno.o o.mom Aoomvma Mo Aoomvme mo m-~u.+mu onsumuwd50a UUQ.+mU musumumdsoe “woo.ov mmo.o o.m- A¢o0.ov mmo.o o.m~m Aao.ov mmo.o o.mw~ Aeoo.ov moo.o o.m¢~ Avoo.ov mvo.o o.mqm Amoo.ov mmo.o m.mm~ awoo.ov moo.o m.mmm Amoo.ov mmo.o o.mm~ “hoo.ov mmo.o m.¢m~ Aaeoo.ov mmo.o m.¢mm Ado.ov oH.o o.mom “moo.cv Hmo.o o.mm~ loomeme s. lommvme 1. oUmH.+mU musumuomeoa mwe mo musuMH0QE0B mumsonHmo msmammoum aw msowumu Edammu vmxmam I500 can mmum mom mmefla coaumxoamm wmum>msmna may no mosmnsmooo musumuwQEmB .vm wanna 148 Table 35. Temperature Dependence of the Transverse Relaxa- tion Times in Propylene Carbonate in the Presence and Absence of C222. CsTPB Temperature * °K T2(sec) 284 0.08 (0.01) 273 0.07 (0.01) 252 0.06 (0.01) 227 0.037 (0.007) Cs+-C222 Temperature * °K T2(sec) 298 0.015 (0.003) 286 0.012 (0.0007) 273 0.006 (0.002) 258 0.0050 (0.0003) 241 0.0028 (0.0002) 230 0.0013 (0.0002) 149 .mcflcopnOua msoucomOEOLCM acumoudou HIuou ma uoonn soaun mosao> ~a\m .musoon ucwxoadeoo msOMun> uo cocoons I Inc was mocummud us» ca cowuuuucoucou +mu Haven 2 «0.0 oanAaucou maceuaHOm uOu h\a MN msoflusc nnHImu 0:» ~00 ma\n mo muoHd manwsom .mn mucosa C no— x 7. 3.. $3551.23. Oh WV 06. m6 I 4 J I 1 d (I I I u I I I 1 q A L or . 1 \ . L 4 .1 . U 100" 1.. i I. I! I H z 000.» \mn 4w 1 VI. 3 . 3 I 1 IV A m-~u.»u . .oooe 150 I 1000: , I CsC222+ b 'I; r O L “J 93. . 1- l ;: 100:- L ~ Cs+ , 1OI' L L 1 1 1 1 1 I 1 1 l l L 1 35 .40 45 (TEMPERATURE °I ¢ x «psue wrem 01* 2. ma »7_ca :0»<4: 4.0 a..« th7uz~3uoxl 2a mz47c0 h-cd Du. :4 ._UJ.m=.m~or. u ch mi» 01 2u<4 .c_mwa «0 ._.~.. mu .hmw> _0Nw~wwmn I pzwrwuwmw.«0.w0n_. u ere—z um» kc w:4<>. p0.u~<~. n puwr mt» h: m: .2..1<>x.x ac ._. _.x m~ I A w w u _ r _ a A U u n . . o — _ ~ 1 u _ m u _ ~ A — _ — _ — .JIIIIVIIIIJIIIIUIIIIT IIIII “IIIIUIIIIVIIIIVIIIIUIIIIUIIIIJIIIIVIIIIVIIIIVIIIIVIIIIUIIIIWIIIIVIIIIVIIII~ > abruo >0 9 42C. »z_co cw»04:uqcu a wrest a .»2_Cd Jdpru1_auoxu 44 vz4w1 x 30.M3mm. u p1u:umuz_.uo.mpn~.I n zoppom 01» »¢ u34<:.:c.u41_. u.OCm wip »« wr44: .2—mwd ac ._.w.x m. 4 .0.mmm~. u ezutmauz_.30.wam_. n exceo 01» re u:0<:.4c.u——.. n bus; 01» p4 u:4¢> .._.o4>x.x do .~._.x vu mvoauc 160 .06 as some 2 Ho.o 0am mesmo z No.o sues wocemuno ssuuomdm 6 no 000 “6056260 .mw wusuem ~3u:~ uc 02w .~omxuu toocnoosoooasooo-onooosu 02w .rIIIIVIIIIVIIIIVIIIIVIIIIVIIIIVIIIIVIIIIUIIIIVIIIIVIIIIwIIIIvIIIIuIIIIwIIIIwIIIIVIIIImIIIIvIIIIrIIII U u x x _ _ _ n x — 7 xx WK — r x . w x x x x x r _ u v n x O O .0 o o o n O s x C C C O D O c C U o 000 COOCC x w u _ 000 x 000 x 0 . 004 o" u x _ u C. O . . .6 yo. x e .V ~ C C x . W ' r O y A c o . ~ ,n. w r w C n 0 v _ _ O _ C T u y u _ xx' _ .u u 1 _ cc 0 _ C V C u C . _ C _ o o p y .u _ x u .0 J O x v _ y p _ o x o x _ .o c _ v m» n v _ 00’ cc — C r _ _ - f u . _ _ h \ _VIIIIWIIIIUIIIISIIIIVII u .u .u v w J m .u .0 m. u u IIImIIIIwIIIIVIIHI . > csrwa >c _ .:c. »:_cc ou»¢4:cs.U0.uro_. u ac» mi» ya w24<> .c_wwa ac ._.N.u m—.; _c.upon. u h:u:uau:_.3c.uo~_. u exceo 01» pa u:4<:.::.uo__. u rum; wt» »4 u:4<> ..~.o<>x.x :0 ._...x v— wvcmu< 161 Table 36. Temperature Dependence of the Exchange Time 1, of some CsC+ complexes in propylene Carbonate and the Corresponding Reciprocal Transverse gelgxation Time of Cs+ (Site A), and CsC+ (Site Temperature 1(msec) —l—(sec) ¥——(sec) °C 2A 28 (A) 18C6 -60.0 3.3 (0.3)b 50.00 58.00 -55.0 2.2 (0.2) 41.00 48.01 -52.0 1.8 (0.1) 40.49 42.99 -44.0 0.59(0.03) 33.00 32.00 (B) DCC -55.0 20. (7.) 43.50 130.55 -47.5 8. (1.) 34.97 92.00 -36.0 3.4 (0.7) 26.50 54.00 -23.0 1.4 (0.2) 20.49 33.00 -10.5 0.73 (0.06) 16.00 20.49 -4.7 0.33 (0.06) 13.70 16.21 (C) C222B —42.0 13.3 (0.70) 30.00 73.00 -36.0 1.8 (0.4) 27.00 55.87 -29.0 0.8 (0.2) 22.99 42.02 -22.7 0.40 (0.06) 20.28 32.47 —10.5 0.08 (0.01) 16.00 20.49 -32.0 1.0 (0.2) 24.51 46.95 a0.02 M CsTPB + 0.01 M b Standard deviation. ligand. 162 Table 37. Temperature Dependence of the Exchange Time in PC in the Presence of C222 and the Corresponding Reciprocal Transverse Relaxation Times of Cs+ (A) and 03+ 0222 (B).6 Temperature 2 1 °C 1(msec)x10 Tr—4sec)-1 -,I.--]-'-—l(sec)-'l 2A ZB 24 1.6 (0.2)b 10.53 68.03 10 4.6 (0.8) 6.08 144.93 0 56. (29.) 13.50 172.41 -12 119. (19) 16.00 178.57 -25 314. (16) 19.01 270.27 -35 523. (19) 21.98 384.62 —37 538. (97) 22.99 429.19 -41 688. (14) 24.51 500.00 a0.02 M CsTPB + 0.01 M 0222 bStandard deviation. 163 T DCC J_J._..LA_LL 5 4 2 ‘9’. 1/ 3. '1 1 r -( ~10F‘ 1 y 4 h 4 * C2228 ‘ J 1 L l J L 1 L l 1 1 4.0 ‘ 3 4.5 (TEMPERATURE 'K )' x 10 Figure 44. Arrhenius of plots of T (exchange rates of Cs+ from ligands) vs T'1 for propylene carbonate solutions. 164 ' I 3 . u: (n 2 F Iv- OJ:- L b “ I p 1 1 1. 1 L 1’1 1 1 4— 4 1 l 1 3.5 4.0 4.5 103 -1 -:r- ( K ) Figure 45. Arrhenius plot of T (exchange time of Cs+ from C222) for propylene carbonate solutions. 165 steric hindrance. The broad temperature range from co- alescence to the slow limit indicates that C222 processes the highest activation energy, but because of the com- plexity of the plot of r 35 (temperature)"1 as shown in Figure 45, the interpretation will be deferred until the next chapter. The activation energies and the exchange times at 298° for the other three ligands were obtained from the Arrhenius equation. Then by using the following relationships: 1‘ = _2_]]<.__ (IV.21) b # = _ AHO Ea RT (IV.22) g AH 7‘ = .. £2 _9 Aso Rinkb Rzn 11 + T (IV.23) AG: = AH: - TAS: (IV.24) all related activation parameters could be calculated (Table 38). In the above equations, AGfi, AH: and A8: are the standard free energy of activation, the standard enthalpy of activation and the standard entropy of activation, respectively, for the decomplexation reaction. The symbols k and h in the above equation represent the Boltzmann constant and Plank's constant, respectively. The activa- tion energy values in Table 38 illustrate that the nature of the ligand significantly influences the complexation 166 o.v« ea m.ow~.~an o.vwav.h o.vuom o.e«o.m moma mo.ouvm.aa o.~nva) m.o«m.h HHHH m.oum.m Adv coo vooo.ouwmb.ma o.~HmHI m.onm.wa mmuvem w.oum.va mNNNU Anomamc maoeos\auo maoaxamox Axommwc maoexflmox gunman mHOE\Hcox o o anomm m. m< m< m m2 8 x 0.3030. .mumconumo wcmaamoum cw moxmameoo owaomoouomz Eaammo meow Souk m0 m0 mmmoamm How muwumsmumm anmcmcoenmna 0cm mmumm mmcwnuxm .mm manna + {Ilif () 167 reaction. Cryptand-2223 has the highest rigidity and also possesses the largest activation energy. In addition, DCC with a substituent, is more rigid than 18C6, and its activa- tion energy is slightly larger. Dibenzo-18C6 has a higher rigidity than DCC but its smaller cavity size and the weaker donor ability of its oxygens makes the complexation with Cs+ much weaker. In addition, it may form exclusive complexes in which the cesium ion is not able to penetrate into the center of the ring. This fact may also be the reason that DBC has the fastest exchange rate at room temperature of all the ligands. If the slow exchange limit for DBC complexes in PC could be observed, it would probably have the lowest activation energy if only exclusive complexes are formed. If this assumption is valid then the activa- tion energy for release of Cs+ ion from crowns would be generally lower than from cryptands. This is in contrast to the trend of the formation constants, which is 18C6 > DCC > DBC m C222 > C2228. It is not difficult to ration- alize the trends caused by rigidity and donor ability of theligand in regard to the complexation reaction. The ligand 18C6 has both flexibility and proper cavity size. Therefore it can form the strongest complexes. 0n the other hand, C2228 has a greater rigidity and, perhaps, the smallest cavity size, and thus forms the weakest complexes. TheAS: values of all three ligands are negative, the magnitude of this value is usually determined 168 by factors such as the reorientation of solvent molecules or conformational changes of the ligand. It seems likely that for the case of C2223 both factors are involved in the transition state. Therefore complexes with C2223 show the largest negative entropies of activation. (B) SOLVENT EFFECT ON THE EXCHANGE PROCESSES Lineshape analysis of exchange rates has also been carried out for the complexes of Cs+ with 18C6 in pyri- dine, and C222 in DMF and acetone. The relevant spectra, collected at various temperatures, are shown in Figures 46-48. For the pyridine case the S/N is poor because of the low concentration of the cesium tetraphenylborate (0.01 M was used) and broad linewidth (m400 Hz) of the resonance line. The nonexchanging T; values for Cs+ in various sol- vents are listed in Tables 39-41, and are shown graphically in Figures 48 and 49. Again, in these solvents the line- widths of Cs+ in the absence of exchange processes are narrow and are dominated by inhomogeniety of the field except at very low temperatures where viscosity broadening starts to become more important. The exchange rates were obtained by fitting the lineshapes with the KINFIT program and the results are listed in Tables 42-44. The Arrhenius plot of log T 2g l/T for the system 18C6 :Cs+ in pyridine is shown in Figure 51. The Arrhenius plots for the cases of C222 in DMF and acetone are complicated (Figures 52,53) 169 Figure 46. 1806 08* :05 IN PYRIDINE Spectra at various temperature for a solution of 0.01 M CsTPB and 0.005 M 18C6 in pyridine. The linewidth at coalescence were N400 Hz. This fact and the low concentration lead to the poor S/N. The dotted lines show the chem- ical shifts of CsC+ and Cs+ in the absence of exchange. 170 A ' -780 C 3906.25 040+ 5: 0 Hz 0222 _ c. -0.5 IN DMF Figure 47. Spectra at various temperatures for a solution containing 0.02 M CsTPB and 0.01 M C222 in DMF. The dotted lines show the chemical shifts of CsC+ and Cs+ in the absence of exchange. Figure 48. 171 W‘ W4 r, :1va WWW) (Wm, 85'er 0 WW. N‘W‘DKMW VNffl/P: .2" W “MW/“WW” “MW“ '37. J ' Csc. C" 0 H2 3905.25 _£c33:3..-_- o 5 m Acevoua Spectra at various temperature for a solution containing 0.02 M CsTPB and 0.01 M C222 in acetone. The dotted lines show the chemical shifts of CsC+ and Cs in the absence of ex- change. 172 Table 39. Temperature Dependence of the Transverse Relaxa- tion Times of Free and 18C6-Complexed Cesium Cations in Pyridine. Temperature Cs¢4B Temperature Cs+-18C6 9C T;(sec) °C T;(sec) -47 0.021 (0.003) -47 0.05 (0.01) -38 0.035 (0.007) ~38 0.07 (0.01) -32 0.05 (0.01) -32 0.07 (0.01) -26 0.059 (0.007) -26 0.087 (0.008) -15 0.087 (0.008) ~15 0.10 (0.01) -7 0.007 (0.01) -7 0.07 (0.01) 9 0.08 (0.01) 9 0.08 (0.01) 23 0.072 (0.008) 23 0.10 (0.01) NOTE: T2(ref) = T2(inh) = 0.1453 sec. 173 Table 40. Temperature Dependence of the Transverse Relaxation Time of Free and Complexed (C222) Cesium Cations in DMF Tempe;ature T;(sec) (A) 03+ 297 0.111 (0.009) 279 0.12 (0.01) 251 0.115 (0.006) 234 0.117 (0.007) 220 0.104 (0.00) 209 0.073 (0.005) (B) Cs+-C222 301 0.03 (0.01) 287.5 0.0230 (0.009) 264.5 0.014 (0.003) 243 0.007 (0.001) 226 0.004 (0.001) 209 0.0019 (0.0005) NOTE. T2 (ref) = = 0.1305 sec. T2(inhomogeneous) 1f74 Table 41. Temperature Dependence of the Transverse Relaxa— (inn Time and Complexed (C222) Cesium Cations in Acetone. Temperature , °K T2(sec) (A) Cs+ 297 0.11 (0.01) 285 0.0 (0.01) 273 0.09 (0.01) 253 0.084 (0.007) 242 0.072 (0.005) 232 0.065 (0.005) 216 0.082 (0.006) 200 0.103 (0.007) (B) Cs*-0222 297 0.067 (0.007) 280 0.048 (0.003) 273 0.043 (0.006) 260 0.03 (0.005) 249 0.040 (0.003) 236 0.031 (0.004) 223 0.028 (0.002) 207 0.020 (0.001) 195 0.015 (0.001) NOTE: - 0.1632 sec. 72(rer) ’ T2mm) 175 .mcwcmomoun msomcmmoeoncw mo muommmm ecu ucmmmummu 00m ca usonm can» mmma mB\H mo mmsHm> .mcwcwumm cw mUmH.+mUIm mo.o no mmamo z mo.o Hmmufim mcficfimucoo mcowu (QHOm How B\H m> anodes: mmalfidwmmo Mom .1 or x 3. 335.35: a T. . 3 06 . ma — a 1 d d —) J 1 u J ~ 1‘ 1. LIL1 B\H mo muoam monEmm .mv madman 0. Z |_13331 ‘3:- OO- 176 Table 42. Temperature Dependence of the Exchange Time in the Presence of 18C6 in Pyridine and the Cor- responding Reciprocal Transverse Relaxation Times of Cs+ (A) and CsC+ (B).a Temperature 1(msec) fi—(sec) Tl—(sec) C 2A 28 -47 4.9 (0.7)b 19.19 _ 47.39 -38 2.6 (0.5) 15.20 28.65 -32 1.4 (0.2) 13.50 21.65 -26 1.0 (0.2) 11.79 16.95 -15 0.4 (0.1) 9.83 11.49 -7 0.4 (0.1) 9.83 14.58 9 0.14 (0.03) 9.83 12.15 23 0.06 (0.01) 9.92 13.86 a0.01 M CsTPB + 0.005 M 1806. bStandard deviation. 177 Table 43. Temperature Dependence of the Exchange Time in the Presence of C222 in Acetone and the Corresponding Reciprocal Transverse Relaxation Times of Cs+ (A) and CsC+ (B).a Tempféature 1(msec) —%;(sec) _%;‘sec) Experiment No. 1 10 2.1 (0.2)b 0.1042 0.0526 0 16 (8) 0.0971 0.0476 ~24 63 (16) 0.769 0.0357 ~37 233 (25) 0.0676 0.0299 ~50 240 (24) 0.0741 0.0244 ~78 1207 (100) 0.1064 0.0147 Experiment No. 2 *24 1.4 (0.2) 0.0300 0.0244 12 2.2 (0.8) 0.1031 0.0541 *0.7 3.4 (0.5) 0.0276 0.022 ~31 107 (17) 0.0718 0.0323 ~41 217 (33) 0.0646 0.0282 ~58 423 (61) 0.0823 0.0213 ~73 1759 (95) 0.1033 0.0161 *Exponential multiplication was used (TC = ~S). a0.02 CsTPB + 0.01 0222. b Standard deviation. 178 Table 44. Temperature Dependence of the Exchange Time in the Presence of C222 in DMF and the Correspond- ing Reciprocal Transverse Relaxation Times of Cs (A) and CsC+ (B).a Temperature 2 1 1 C I(msec)x10 §—-(sec) fi——(sec) 2A 28 ~8 0.26(o.02)b 0.116 0.0133 ~12 0.61 (0.07) 0.1163 0.0119 ~20 0.8 (0.2) 0.116 0.0095 ~24 1.31 (0.35) 0.1163 0.0085 *~31 1.8 (0.1) 0.116 0.0056 ~37 4.9 (0.9) 0.1163 0.0057 *~55 44 (9) 0.0290 0.0086 ~67 288 (33) 0.0606 0.0019 ~78 1944 (123) 0.0294 0.001 *Exponential multiplication was used (TC = ~5) a0.02 M CsTPB + 0.01 M 0222. bStandard deviation. 179 .mcwcmcmoua msomcmeEoncw ucmmmummu owm 0H usonm can» mama mB\H mo mwsHm> .mcoumom cw use use cw +~mmomplm mo.o Ho mmamo z mo.o Hmcufim mcficwmucoo mcowusaom now B\H m> anodes: mmalesflmmo 0:» How ~9\H mo muoam moHHEmm .om musmwm as no. x 3.6 $23.13: as 3 3. mm m.-.-‘...‘ 9-00.00-00-0‘0900“ * 4% o— -‘--m‘. ‘000 .‘*“O L ”0‘08‘0888 * L / ‘0h‘88“““““““ m H ‘*“|\|“' U - Loo— l+IAJA coo— III In!" l..||ln 180 0.00! U I I#' T 1'ISECI "' O-OOOI lj f I '— " 0.0°°OI r l 1_ L j A l l ’1 l j 35 Figure 51. 4.0 _ 45 l TEMPERATURE °Kl 1x103 Arrhenius plots of T (exchange time for Cs+ ion from 18C6) in pyridine, 181 I I I I T IO lj I—I'W'I '? A P U )- w )- (n . E . k . OJ ' Uti‘fl r 1 I 1 1 1 1 J 1 1 1 J 1 L4 1 1 l 1 . 3.5 4.0 4.5 5.0 3 -: J— K) T ( Figure 52. Arrhenius plots of T (exchange time for Cs+ ion from the ligand C222) in acetone solution. 182 IO:- )- '1— A : o h u: . co :5 . t- * OJ :- ' I 0.0!:- 1 1 1 l 1 1 1 1 1 1 1 L 1 L 1 1 4-0 4.5 5.0 3 1o -1 T Figure 53. Arrhenius plot of T (exchange time for Cs+ ion from C222) in DMF solution. 183 and will be discussed in the next chapter. The activation parameters for the release of Cs+ by 18C6 in PC and pyridine are listed in Table 45. The Ea values for both cases are similar ; however, the big stan- dard deviations make the meaning of the difference ques- tionable. Previously, Cahen had observed (6) in the Li- cryptate study that the activation energy increases with increasing donicity, Opposite to the overall energy change. Therefore, he concluded that the transition state must in- volve substantial ionic solvation. In the present study such an effect seems to be small. (C) DISCUSSION Shporer and coworkers (4,5) studied the dynamics of complexation of Na+ by DCC in DMF and methanol and reported Ea values of 8.3 :1. (Kcal/mole) in both solvents, while for the complexation of Na+ or K+ with DBC in methanol the Ea values are 12.6 i l. (Kcal/mole) for both cations. Therefore, they concluded that the complexation reaction of Na+ with DBC has a larger activation energy than with DCC because of the higher steric hinderance of DBC, they also concluded that there is no solvent effect since the activation energies for Na+ complexes with DCC in MeOH happen to be the same. The value of Ba found by Shporer and coworkers for Na+~DCC in methanol is about the same as we have found for Cs+~18C6 in pyridine. However this 184 Table 45. Exchange Rates and Thermodynamic Parameters of 18C6 Complex Exchange in Propylene Carbonate and Pyridine. Activation Parameters PC Pyridine Ea (Kcal/mole) 8.0:4.0 8.04:0.27 kbx1o'3 50.014.o 8.9 20.8 at 298°K (sec‘l) g AHO (Kcal/mole) 7.41:4.0 7.5i0.3 As: (cal/ K mole) ~12.2 $0.2 ~15.5i0.2 AG# 0 (Kcal/mole) 298°K 14 14.0 12.1:0.4 185 could be coincidental because of the compensation of several factors which affect the activation energy. For instance, 18C6 has less rigidity in the ring than does DCC, but a larger cation such as Cs+ might introduce a higher activation energy. Therefore, a coincidence could occur by compensa- tion of the ligand effect and cation effect and solvent effect. In other words, the factors which affect the reaction rate are complex, and conclusions cannot easily be based upon comparisons made with very different systems. Together the solvent effect and the ligand effect of the complexation reaction can be represented by the reaction profile drawn below: I” ‘\ W I, ‘ \ ,’ \ , 1 I I 1 0" /’-.\\ 1 EC! . I I I ’9- ‘ f,’ I \1 X ,' I ’ \\ 1 I I I’ 11 1 + 1”, \\ “ C5 (PC) '12:-J’ ’_ “ ‘\ \‘ E0 ’1’ “ \ \ T- Ea CJWPY) _,,.’ 1 ‘\ \ \ |‘ \\ \ Ea ‘( \ \ , + \\ \\ CsC2228 (PC) \ “\ \ 1 + \\ ‘ Cs DCC(AHPC) \ ‘ * Cs+18C6 (pc) \ X r 031806 (PY) CHAPTER V A CESIUM-133 NMR STUDY OF THE COMPLEXATION OF CESIUM TETRAPHENYLBORATE BY C222: DISCUSSION OF THE TYPES OF COMPLEXES 186 I. INTRODUCTION The diameter of the cesium ion (3.3 A) is larger than the estimated diameter of the C222 cavity (2.8 A). How- ever, an X-ray study of the solid complex (55) showed that the cesium ion can be accommodated into the cavity of C222. In the present chapter the discussion will focus on the nature of binding of Cs+ to C222 in solution. II. EVIDENCE FOR TWO TYPES OF CRYPTATE COMPLEXES Inclusive cryptate complexes were first designated by Lehn (50) with the mathematical symbol (2i, M+C L). It was considered that whenever the cavity diameter of the cryptand (L) is larger than or equal to the diameter of the metal ion (M+), inclusive complexes are formed in which the cation is situated within the cavity. However, in the present discussion, we need to include other pos- sibilities. The symbol M+C: L, will be used to designate a complex in which M+ is located entirely within the cavity (inclusive complex). By contrast, whenever a complex is formed in which the metal ion only partially penetrates into the cryptand cavity it will be called an exclusive complex and designated as M+lJ L. It has already been mentioned in previous chapters that the reaction of Cs+ with [2]~cryptands seems to give more than one type of complex. A summary of the observations 187 188 which support this speculation are as follows: (1) The limiting chemical shift of Cs+~C222 (measured directly at large ligand/Cs+ ratio or obtained as a parameter from data fitting) is very solvent dependent at room temperature as shown below. Solvent PC Acetone MeCN DMF DMSO Pyridine (Appm)lim -l93.8 -203.0 ~210.4 -155.8 ~144.3 -224.3 (2) The limiting chemical shifts are also temperature dependent, as shown in Figure 54. A special feature is that the extent of variation with temperature de- creases at low temperatures eventually extrapolating to the same Cs+ chemical shift in all solvents. By contrast Cs+-(18C6)2 complexes for example, show neither solvent dependence nor temperature dependence of the chemical shift. We would expect Cs+ inside of a C222 cavity to be well-shielded from the influence of the solvent. (3) The spectra obtained with samples which have only enough C222 to complex half of the Cs+ (1:2; 0222/ (Cs+)t = 0.5) in five solvents were collected at various temperatures including temperatures low enough to be at the slow limit of the exchange process. As shown in Figure 55, the chemical shifts for the complexed + Cs at low temperatues have about the same value. However, in PC and in DMF solutions the linewidth of 189 .msHm> menu on» on moum>coo mucm>H0m omfimw Haw How Amocwa owuuocv mcowumaommuuxo 0:9 .musumumafiou m> mcouoom can om .mza a“ -~o+mo mo uuaam Hmofismao sceufisfia mo scan a .qm musmflm co_wmah Hmucoefluomxo 00008 x was 00am» cmuma .00H0> Hmucmeanmmxmatfl3 Axov musuwuomfimu m> on c« (seam 00008 0 mumaz ~m~0+mo mo numecm Hmowfimno mcwquHH «0 uoam cupcasfiwm mo camflumgfioo é .mm .60: 335.35.. mmm mmm GNP fl - q a 4 000+ 000 0 0* «a . mac 0 00% Q 00E 000 O O L 000 000 0000 000000 whamwh OONI wddV cowl 198 solvation between the complexed ion, the free salt and the free ligand, ASC is the net entropy change due to con- formational changes, and ASn is the change in entropy caused by a change in the total number of particles. In Cs+~[2]cryptate complexation, ASn should be nearly indepen~ dent of the type of complex formed, and it will be left out of the present discussion. The term A88 is probably positive because of the displacement of the well-ordered solvation shell of the cation (except for a highly struc- tured solvent). The ASc term is likely to be negative be- cause of the decrease in the flexibility of the cryptand when it forms a complex. According to these arguments, we expect AS to be negative for the conversion of the exclusive to the inclusive complex. This speculation fits the esti- mated entropy changes given in Table 48. Table 48. Estimated Entropy Values for the Complexation Reaction at 298°K. Solvent Asl(e.u.) ASZ(e.u.) AS(e.u.) PC -l3.7 -7.0 -20.7 DMF -11.2 -7.6 -18.8 199 It is interesting to note that A81 is strongly solvent dependent, but that A82 is much less dependent on the sol~ vent. The existence of two types of complexes in the solu~ tion would also be expected to affect the apparent activa~ tion energy. Indeed, we obtained anomalous Arrhenius plots as shown in Figures 45 and 52. The curved Arrhenius plots for the acetone and PC cases could result if the exchange process involves more than two sites. If we assume that the reaction actually occurs in two steps as given by Equation V.l then the activation energy for dissociation of the ex~ clusive complex must be higher than that for the inter- conversion of the two complexes. This phenomenon is shown by a collapse of the single line at low temperatures into two lines, one of which occurs at the chemical shift value of the free ion. If at low temperatures, the second re- action becomes slow, then a broadening of the line for the complex will occur which will give an anomalously large value of T when only the two-site exchange process is considered. It is simply not valid to use the two- site expression when the rate of conversion to a third site becomes slow. This speculation is also reinforced by the exchange spectra of Cs+C222 in PC and acetone at various temperatures shown in Figures 37 and 48. It is interesting to note that at low temperatures when the exchange rate between free and complexed ions is slow enough to permit observation of 200 two distinct peaks as the temperature is lowered, the com- plexed cesium peak is first broad, then narrow and finally again becomes broad. This effect is especially pronounced in PC solutions. The broadening of the signal for the complexed Cs+ species could be caused by a slowing down of the exchange equilibrium, [Cs+L)C] : [Cs+CZC] at low tempera- tures. The lineshape analysis for DMF solutions at various temperature also shows a broad Cs+C222 line at low tempera- ture. However, it was found that the exchange process had not yet reached the slow exchange limit. This may be the reason why the Cs+C222 kinetics study in DMF gave a straight-line Arrhenius plot while it was curved in the other two solvents. The activation parameters for the ex- change reaction in DMF were therefore calculated. Table 49. Activation Parameters for the Release of Cs+ from Cs+C222 in DMF E = 13.520.3 Kcal/mole a. _ 6 -1 (kb)298 -(910.9)x10 sec (10%) = 7.8:0.6 Kcal/mole 298 AHg = 12.910.3 Kcal/mole # A50 1721 (e.u.) CHAPTER VI SUMMARY AND SUGGESTIONS FOR FUTURE STUDIES 201 I . SUMMARY Chemical shifts of the cesium—133 nucleus were measured in six nonaqueous solvents relative to 0.5 M aqueous cesium bromide. Cesium tetraphenylborate (CsTPB), triiodide and thiocynate were used to determine the infinite dilution chemical shifts in pyridine (PY), propylene carbonate (PC), dimethylformamide (DMF), dimethylsulfoxide (DMSO), aceto- nitrile (MeCN), and acetone. The corresponding ion—pair formation constants were determined from chemical shift- concentration data with the aid of a weighted non-linear least-squares program (KINFIT). The association constant for CsSCN in pyridine is 9002200 while for CsTPB in pyri- dine it is 370120, in PC it is 1617, in MeCN it is 40:10 and in acetone it is 2213. The uncertainties given are standard deviation estimates. Cesium-133 NMR studies were also performed on CsTPB complexes with five ligands in six nonaqueous solvents mentioned. These ligands were 18~Crown~6 (18C6), dibenzo- 18C6 (DBC), dicyclohexyl~18C6 (DCC), cryptand-222 (C222) and monobenzo-C222 (C2223). These ligands have different topologies and substituents which affect the complexation ability. Cesium tetraphenylborate forms both 1:1 and 2:1 (1igand/Cs+) complexes with 18C6 in all six solvents with the first formation constant (K1) larger than 103. A new EQN subroutine of the KINFIT program was written to analyze data which show both 1:1 and sandwich complex (2:1) formation. 202 203 It was found that both K1 and K2 are affected by the geom- etry and substituents of the crown ligands. It was also shown that the solvent plays an important role in the equilibrium process. For example, the K values for 18C6 in six solvents are in pyridine K1 > 108, K2 = 7111: in PC x1 = (1.510.6) x 10‘, K2 = 8:2; in acetone K1 > 107, K2 = 3410.5; in DMF K1 = (913) x 103, K2 = 2.44:0.05: in DMSO x1 = (1.1:0.1) x 103, K2 = (110.4); in MeCN K1 > 105, K2 = 4.4:0.3. When substituents are added to the ring of 18C6, the resulting differences in geometry and rigidity affect the K values substantially. The complexation reaction of all three crowns in pyridine can serve as an example, the ligand 18C6 is the most flexible of the three crowns and has the highest K1 value, while DBC with the smallest cavity and greatest rigidity has the lowest K1 value. On the other hand, because of its small cavity and high rigidity, DBC may form a non-symmetric 1:1 complex and this geometry favor the formation of a 2:1 "sandwich" complex. In any event, DBC gives the highest K2 value of the three crowns. Dicyclohexyl~18C6 has a similar cavity size to 18C6 but a higher rigidity of the ring. It has an intermediate K1 value, but the smallest K2 value. A thorough study of 18C6 complexes with CsTPB in pyri- dine was made at various temperatures (from 25° to ~44°C). For the purpose of this study, a new temperature independent reference for all temperatures was designed. Its validity was tested and provided evidence that ion-ion and ion-solvent 204 interactions are temperature dependent. In this complexation study, conclusive evidence for the 2:1 complex (18C6:Cs+) was also found. The values of the first formation constant at various temperatures were too large to be determined by NMR techniques, but K2 values were determined and used to obtain the enthalpy and entropy changes for the second com~ plexation step. The results are: AH = ~6.210.1 Kcal/mole 2 (AG -2.83i0.004 Kcal/mole, A82 = -11.210.3 e.u. o) = 2 298 A kinetics study of the decomplexation reaction of Cs+o18C6 gave an activation energy of 820.3 Kcal/mole. The formation constants of C222 and C2223 complexes were also obtained from the NMR chemical shift data. The formation constants for these two ligands showed the same trends with various solvents. For example, the K1 values 5 3 for 0222 are > 10 (pr), (1011) x 10 (PC),(10.8:0.8) x 103 (acetone), (1.510.1) x 102 (DMF), (27:3) (DMSO), and (411) x 104 (MeCN). By contrast, the 02223, with a benzo group on one of the ether chains forms weaker complexes with K values ranging from (S.7:0.8) x 103 (PY) to zero (DMSO). Chemical shift-mole ratio temperature dependent studies were also carried out, as well as studies of kinetics. Both gave evidence for two types of complexes in the solu~ tion. The results are interpreted on the basis of the formation of both inclusive and exclusive complexes of Cs+ by C222. Enthalpies and entropies of formation were com~ puted by using the KINFIT program and it was found that both quantities are sensitive to the solvent for the 205 complexation of free cesium ions to form the exclusive complex. The conversion of the exclusive to the inclusive complex is much less sensitive to solvent. The activation energy (Ea) for the loss of Cs+ from the Cs+C2223 complex in PC is 14.910.6 Kcal/mole. By comparison, the values obtained from kinetics studies for crown complexes give Ea = 8.5:0.5 (Kcal/mole) for DCC complexes and Ba = 8:4 (Kcal/mole), for 18C6 complexes. It appears that the higher rigidity and special geometry of the ligand give C2223 the largest activation energy. II. SUGGESTIONS FOR FUTURE STUDIES The studies already made lead to the following sugges- tions for further investigation: (1) Calorimetric studies of the complexation reaction can supply some complementary information to the alkali NMR results. A joint study of Cs+~C222 in various sol~ vents by calorimetric and NMR lineshape analysis should give more insight into the nature of the complexation reaction. (2) It has already been observed that the chemical shift of cesium ion in electrolyte solutions is tempera- ture dependent. By using a temperature independent reference, a series of studies such as Kip for an electro- lyte solution and formation constants for complexation 206 reactions could be made at various temperature and thus one could obtain the corresponding thermodynamic parameters. Also, it may be worthwhile to investi- gate ion-ion and ion-solvent interactions at different temperatures. (3) At present a few studies have been made of the Cs+ counter ion effect on formation constants. Therefore, a complete study of cesium salts complexed by [2]- cryptand could be carried out. This information could help to determine the conditions which favor inclusive Kg exclusive complexes. It could also be used for the study of competitive complexation reactions such as that between Cs+~[2]cryptates and Na+~[2]cryptates. Since the latter complexes are usually stronger than those with Cs+ the extent of replacement of Cs+ by Na+ could be used to obtain relative complexation formation constants. (4) Preliminary temperature studies of the C2223 complexation reaction with Cs+ in PC and acetone seem to indicate that the chemical shift of complexed Cs+ is not very temperature dependent. Therefore, a thorough temperature study of the Cs+~C2223 complexation reaction in various solvents may give further information about a system which probably forms only exclusive complexes. (5) A lineshape analysis at various temperatures and 207 concentrations could be made of Cs+-[2]cryptate complexa- tion reactions, to determine the order of the decomplexa- tion reaction in low dielectric media. It has long been of interest to find out whether the mechanism of decomplexation involves a unimolecular or a bimolecular process. If C221 or C2223 form only exclusive complexes with the Cs+ ion, this condition may offer a very favorable situation to test the nature of decomplexa- tion mechanism. APPENDICES APPENDIX A DETERMINATION OF ION-PAIR FORMATION CONSTANTS BY THE NMR TECHNIQUE; DESCRIPTION OF THE COMPUTER PROGRAM KINFIT AND SUBROUTINE EQN The equilibrium for an ion pair reaction can be expressed as and ip = I c in which Kc is the concentration equilibrium constant and y+ is the mean activity coefficient. By using the well known Debye-Hfickel equation, y+ can be thus calculated as follows: 1.823x106 | |/_ 3/2 2 z_~ I .109 7+ = ‘DT’ + (A.1) ’ 50.29 1 + —————7— 3 /I In this equation 2+,Z_ are the charges of the ions, I is i (C = Concentra— the molar ionicstrength which is 1/2 ZCiz tion summed over all species in the sglution), D is the dielectric constant of the solvent and T and 3 are the temperature (°K) and the closest distance of approach of the ions in A. 208 209 The observed chemical shift is a p0pulation average of those of the free ion and the ion pair; i.e. obs 5 EXP + 6ipX ip )XF + 6. . (A.2) = (GFG ip 1p where ME = [Cs+]/C¥; and Cg is the total concentration of Cs+ in the system. Material balance gives 0: = [03*] + [Cs+ - x’] = [Cs+] + Kc [Cs+]2 Therefore M ~ 1 1+4K C [CS+]= J C T ZKC + x =C3 =-+J1+4KC 0?; 2 F EM M and KC = Kipyt T 2K C C T So that, finally -1 + (1+4K. ch yf)1/2 Gobs = (6F ~ 61p) + 61p (A.3) 2K CM ' 72 ip T 1 In order to make computation easier on the CDC 6500 com~ puter, Equation (A.l) was rewritten as: 210 1 6 1.823x10 _ 2.303 |z+z_|/f ‘ 50.29 1 [(DT)1/2]3[1 + 73;;172 Eff]J -791600|z+z_|/i exP 1/2 3 50.29 [(DT) ] [1+—__—I77 Slf] (DT) Two unknown parameters must be obtained from the fitting procedure. These are the limiting chemical shift of the ion-pair (U(l)) and the activity equilibrium constant (0(2)). The known parameters are introduced as constants; i.e.; const(l) chemical shift of the free ion, const(2) = (dielectric constant) x (temperature) and const(3) = average distance parameter 3. In this study a value was chosen from reference 100. The Fortran eXpression and deck struc- ture are listed on the next page. It should be noted that in these cases, all K T M general a correction should be made by considering the degree ip values were less than 400; therefore, C was used to calculate the ionic strength. However, in of dissociation. 21J1 0.0.0.000...00.00.0000...’OOQQOOQOOOOO0003cOOOOOOvoOOOOOOOOOOOOOIOOOOIOO ' IS? of [0V PAIN FGHMATInN cons IAN! nglsnutwartnu 0 0000.005... OOIOOOOOOOOOODO.CQ “once-anonuoeooool OOOOOOOQQOOOOOOOOOOOO ch J08 cAnn PISS H090 HALOI'UYEoKlN‘Itfio 8 9 EHRSOHIINE rflfl OMHON Kouuf. oIYADEoIVADFOIUTOIAnalINCQoNOpT.NOVAQQNOUNéoXoUOlTHAXoK luTx.1th.l.Av.Q£s|n.Ino.ip§.lfvp.xx.oxvvp.nx1I.FoDoFo. m.poZL.InoEK PIGV‘LoXST Y'nToLoMsl"oYonYoVECY NFS I CONSIQNDAYvJU‘To“0°ToLn° To 3vvv.COnSgS COMMON/F COHVON/DOIN nluFNq 0N u 0 R ‘u 300) o ( l ) X 90) 3".“ FOPIJOOIor0(JOO)oF #(20):XST(J00).Y(I S:P(50)00XII(50)91 U- :30 - nFHWWQfiDZZFWtL-«530 I.-. g '3' II"? VV-)):~fi° Q-WNWVOCH’ (fidDDZ ”E MINATION 0F K(IP) HY CS-NNR ' ) ) bizp¢uo-ro— .- b - — m " E IOZV onvvzonm mevanuunnn xaotolnb cccccccccc FREE ION fl 0 n .. 0"D’MM5DF) on .C CCCCCCCCCCcro3ccrrtr CCCC ”(:13 ACIIVIIY Inn DA! - b .— ~ - V C C C C C CCCCCC )o unwaqnnum—r—dfiaonxnumnm—mrmoo fmfiDOCVMMWfiOmOFKK7304-NLDInmmx-n«:«cq OCWNWO foJVD orvn 2nd- J‘nfiix «nt- Anal. OP was! on .0 WC) On an Fr} DC) CCCCCCCCCCCCC C 050.290CONSY 3 ‘u-IXJFUNWJ .aOthntbh ‘ 00—0 ODOD—nfi ‘3' .fi —o-n ""'fiffimfib \th~ — OU‘NVfiZ o-«muo:\ Z-JWDZ .nxfiZJ) HXCWWfi «ACWWD V—f) Q 3 /(Z.OCONKOXX(1)))O(C0NSY(1)-U(I)DOU(I) v-uJWA—FHN'DZP) -D>h—D~u1<--—n quXSMHOCWWV-GC) 3% «24:4n-§ 'dC-“:4C5‘C-ao DFHNWDOINDfi‘WV v MONOJIDW M II n O 20 C MIINUF lo CO~TI~UE 1| CONIINUE RETU 17 CENIINUE R TUR END 1 n .IOOOCQOOOOI.DOOOOOIOOOOCOO.OOOOOOOOO9°.3COoI0.0.0OOOOIOOOOOOCOOOOOOOO.. CONTOOL CARD onunN capo ~ ONST NTS INITIAL ESTIMAYE par: cnon ooooooooodooooooobooo0.0099000060000900:-voooooooooooooooooooooooooooooo zLANv CADO ? H 9 212 APPENDIX B DETERMINATION OF COMPLEX FORMATION CONSTANTS BY THE NMR TECHNIQUE; DESCRIPTION OF COMPUTER PROGRAM KINFIT AND SUBROUTINE EQN (l) DETERMINATION OF FORMATION CONSTANTS FOR A SINGLE STEP REACTION The equilibrium for a one-to-one complexation reaction can be expressed as M + L I ML (8.1) and the concentration formation constant K will be K = CML/CM‘CL (13.2) where C stands for concentration in terms of molarity. In this case we only considered the concentration formation constant since there are no changes in the number of charged species and the concentrations of the reactants are very low (0.01 M). Therefore, we can consider that the mean activity coefficient for the free salt and the complexed cesium are the same. The method used in this technique makes use of the fact that at the fast exchange limit (on the NMR time scale). the observed chemical shift of M (Gobs) is a weighted aver- age of the characteristic chemical shift of M at each site 213 (free M and complexed M in bulk solution). 6 = X 6 + X 6 (3.3) where 6M, 6 are the chemical shift for free and complexed ML M, respectively in a given solvent, and XM' XML are the relative mole fractions for each species. Therefore obs _ M = XM(5M - 6ML) + GML (3.4) The analytical concentration of M is CE = CM + C and for . . T _ the ligand 18 CL — CML + CL ML _ T _ T _ T so CML — CM CM CL - CL CM + CM so _ T T T Then K - (CM CM)/(CM)(CL CM + CM) 2 T T T _ K CM + (K CL - KCM + 1) CM - CM - 0 _ 1 _ T _ T f”“T_ T *2 If CM - if“ { (K CL K cM + 1) i. (K CL K CM+1) +4K CM } (8.5) Since physically C cannot be negative, therefore only the M positive root is chosen. Let T D = (K CL K C Then 214 + 1)2 38 __ 1 T__ T ‘E—‘T cM - if {—(K CL K cM+1) + +4K cM } and 5 = 1 {- obs T 2K CM (K CE—K C13”) + \lD-MK C§}(6M - 5 + 5 ML) ML (8.6) In order to fit this equation, two constants and two param- eters are used; namely: u(l) = GML const(l) = C: The above equations follows: - T = A - K CM u(2) _ T = B — K CL u(2) C _ (GM-5m) = T 2K CM u(2) = K const(2) = 6M expressed in Fortran notation are as * const(l) * xx(l) (const(2) - u(l))/(2.*A) D = (B - A + 1.) ** 2 S = 6 = ((A - B - l.) + SORT (D + 4.*A))*C + u(l) obs 21J5 cocoooonouaoonoonanouoeooov "n sovnomur:oao.11--:1a1o:-ua<.w an rrlrét‘ooc-nnncunnoo3.1000 .; 0.0.0 ((PNP nfrfiwulNATlnN nF rnRMAtjnN CanTA~T o ogsqoopoqoooonnoOcooe1cnooo 74 sxnuuz~~wenwcvengeanooooonuooooooo "El.ruasnnfl.T)dr§Rq?.Jr«a.. DU=QCCQFTCTHC. HAL.“ HQLQI- 7 A o ANMFGoDUF CP’Z. 7's ‘. onyr.xTuF[7|. SHnCJnTTNF FUN Cfluunm KOHNI [TADF.JTAJrql.ToLAH.T[1C).JHUToNHVAKoNOHH.a..xcx..o__ qu:o A1o.c.u.c.o~u.a.o_u.o.oFu.xcmc~c parccu UH“ .m.~nx.m.,xv>.<..zvxc.m__ ht—QB .x.>_n.xc> u“ .xcxculxcx yv<2.1x.>~I,xc>L.moooocom. mo..z.>—cu~ m. ~ux m- cc .o_cm.xpv hazccu c- .~._u~..u.>_.._.x~..o_~ pzcca m . .opcmc hazucu so“ m.m.._:nzL_u.u0mc L. .m._uu.._.>_.._cy~..mo~ oqwa c~ .m.o.um. eczocu co~ o..ooo..»:n2~4m.urmc up x.<.oo. aqua x mthn >m-~hhwnphsh~han4zvckcc .A.>. Nox.,m.>_..m~x~ zcuvzuzwc .mcnILzzo. mcu L;_cp :.u¢ :cz—.h&u>zcc :caLcoo o a w .cc. .2pu .wwz.cw7274" 3n .ocu o_.cccmo.n co. at». occmcrL.~w: L2d . ssasstssssssaattscvvstsas¢¢¢$t$¢a¢$$$$$¢$$$$$$¢¢¢¢cessattnvtanactttssatc a ohmm>ZCU Idaocaa uC hU—J B $3avfififififiavfivfivt.4$33.33¢*#$$$¢¢$$.¥$$$¢.3«a.u..n.a..n...._..$*8.$w..#.¥#$$fl.$fi.$$$##fi.fi.¢$$#fl..¥¢¢$$$$¢ 233 0 MnanrHANfiE CASI LINCSHAPE ANALYSIS IOOOIOOIII00......DID.O.000.00....OOOOOIOOOOIOOOOOOIOOOIIOOOOOIOIOOIOOI. IIPNC PIGTIS 073|76 07 DYE MEIICMasooo. TIOOcR 67oJC10 . PH=SECRETSTHF. gALI .L'DYEQKINF‘TIO s 9 sunnnHIINE Fun cnuunn KOUNI:l[ADFoJIAHFOIWII5A90XINCHIMOPTcNOVAQB Mnuuxox.u.lrunxcrnu ’wtx.trqtol.av.wffilno[ADIEPSoI YpoxxowaYP.nKIl.F09 UFO FU.DoZLoIOoEFON .IGV‘ o‘S'Ot'ntOLoMIJ.JoYonYOVECIQNcgtoanQY DIMFNSION xta.loaI- "I70I. waI«.IO0I. XXIhIo PODIIOOI o FnIIOOI. FFnN I285 IuIIooI. PIZCo’lIo VErYIROIZII. ZLIIOOI. 10(20Io EIG VAL IZOI. XSIIIOEON I286 ZOI. YIIOII DYIISI. CnNQIIlbI _ GO TO (70304990|Io I'VP I‘CONIINHE IIADF=6O J7APE8h| HRIIE IJIAPE AI C rnnuan 0 or“ An Ic qvII. VECII7oo?IIp ZLIIOOI. 10(70Io EIGV‘L‘ZOI' XSIIIOEQN I286 20). YIIOII UYIIOI CONSIII6I . GO IO (79395959lIo [IVD [ an»! NHF ITAP :60 .ITAF sfll . WIIE. IJTAPEoSI _ 6 rnnuAiI 0 DYNAMIC SYSTEM or COMPLExAtInN 'I Nnvanaz RE11?N E CCCCCCFCCCCFCCPCFCCCFCrrrrICrrCCCCCCCCCCFCCFCCCCCCCPPCCCCCCCCCCCCCCCCCCC c HIII=K U(?I=C Ht1I=IIU u(«IatHEIA A UISI=FREO 0'7 SET. c 2 CnNSEIAI=v:AIC Cnu:z::I31PéA. CIIQI‘E‘APIH’C C"?§§{“’='7‘°' c I =W n I =N| ' H T 8 .D I” I géADAH/MEHOHIVW :9 E ONSII ONS N0 1:: cancer ¥ESSC5$5E$S$SCCCCC .Ccé CCCCCCCCCCCCCCCCCCCCCCCCCCCCCCCCCCCCCCCCL gs CONSIIII/CUNSII7I 0 CnuSIIII/COHSIIhIOUIII/MCONSTI7I'CONSIIQI I I -HI1I0ICnNSII