l I l v MY, 1”“ I 5' H“; I -‘ v -- “I I : " a I MI” W I' .’ ‘I III: ‘I‘ t .l \. “ It‘ll!!! 1,1 I IVW’AAI‘II\I :w‘fillflhfl'h‘l‘nujm’ I. Dl‘l I. 4 I I f‘l : . 14 r I: I??? IL Hill: I III}!lllllllzllllfllllLlllllljllllllllllljlll F iii; :2 A R 2" This is to certify that the thesis entitled HoH'fi-wc\eOY N H22 S+ud~l o; (ompwafion of SOY“ DV‘RVCXLQAA“ Ccd’iOYtS‘ b7 Crown ETRCKS “ Monotguous Solmls. presented by Hoj’i’Q ba Shot/m Sipu r has been accepted towards fulfillment of the requirements for 911 D degreeinm- h'7 Major professor Date /' 30'77 . ., fix ., .4 OVERDUE FINES: , ‘ 25¢ per day per item RETURNING LIBRARY MATERIALS: __________._.—-——-—- “ ”(rm-xx 1" l“ v.1 ‘93". \"“VW $5]; ' Place in book return to remove ' charge from circulation records MULTINUCLEAR NMR STUDY OF COMPLEXATION OF SOME UNIVALENT CATIONS BY CROWN ETHERS IN NONAQUEOUS SOLVENTS By Mojtaba Shamsipur A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Chemistry 1979 ~r. L. : s S C . a. r” hug Q» h. 7‘ Q. o“ «C C 3 AC 7.. .3 g. v e 1“ ab a: Zr «0 A: é/Agsrk ABSTRACT MULTINUCLEAR NMR STUDY OF COMPLEXATION OF SOME UNIVALENT CATIONS BY CROWN ETHERS IN NONAQUEOUS SOLVENTS By Mojtaba Shamsipur l3C 23Na, 13305, and Nuclear magnetic resonance of , 205Tl nuclei were used to study the sodium, potassium, cesium, and thallium(I) ion complexes with dibenzo—BO- crown-10 (DBBOClO), dibenzo-ZM—crown-8 (DBZNC8), and dibenzo- 21-crown-7 (DB2lC7) in nitromethane, acetonitrile, acetone, methanol, dimethylformamide, dimethylsulfoxide, and pyri- dine solutions. With the exception of the DB30010-Na+ system, all of the complexes were formed in 1:1 mole ratios. The presence of three sodium DB3OClO complexes, Na2(DB3OClO), Na3(DB3OClO)2, and Na(DB3OClO) in nitromethane and acetonitrile solutions was deduced from the behavior of the 23Na chemical shift as a function of DB3OClO/Na+ mole ratio. The NMR data support the existence of a "wrap around" structure for the DBBOClO complexes with cesium and potassium ions in solution. n3 «\g "Y‘. 3» a: a w . 1 u‘l‘ Q» \~q Shir. h. n . .f Mojtaba Shamsipur The stabilities of the sodium, cesium, and thallium(I) ion complexes with both DB2HC8 and DB21C7 decrease in the order Tl+-Crown > Cs+-Crown > Na+-Crown in solvents of low and medium donicities such as nitromethane, acetonitrile, acetone, and methanol. The ability of the ligands for the formation of a three-dimensional "wrap around" complex with the same cation decreases with decreasing the size of the ligand, liéi: DB3OC10 > DBZUCB > DB2lC7. In all cases studied, with the exception of pyridine, there is an ex- pected inverse relationship between the donor strength of the solvents and the stability of the complexes. The extent of the solvent effect on the complex formation +, and T1+ ion complexes de- + constants for the Na+, Cs + > Cs creases in the order Tl+ > Na The chemical shift of the 133Cs resonance was studied as a function of the ligand/Cs+ ion mole ratio at various temperatures in five solvents, iiii: nitromethane, aceto- nitrile, acetone, methanol, and pyridine. From the result- ing data AG°, AH°, and AS° values for the complexation reactions between the cesium ion and DB3OClO, DB2NCB, and DB21C7 were calculated. It was found that in all cases, while the stabilities (or the AG° values) of the complexes are not very sensitive to the solvent, the enthalpy and the entropy values vary very significantly with the solvent. In all cases the complexes are enthalpy stabilized but entropy destabilized. From the results, it seems Mojtaba Shamsipur reasonable to assume that the main reason for the negative entropy of complexation is the decrease in the conforma- tional entrOpy of the ligands upon the formation of a metal complex. +, Cs+, and Tl+ ions by 1,10- The complexation of Li+, Na diaza-l8-crown-6 (DA1806) in several nonaqueous solvents was studied by multinuclear NMR technique. The formation constants of the resulting 1:1 complexes were calculated by computer fitting of the mole ratio data. The stabilities of the complexes decrease in the order DAl8C6-T1+ > DAl8C6-Li+ > DA1806-Na+ > DAl8C6°Cs+. In order to study the effect of the substitution of two nitrogen atoms for the two oxygen atoms in lB—crown-6, the formation constants were compared with those reported for the l8—crown-6 com- plexes with the same cations. As expected, the sodium and the cesium ion complexes are weakened appreciably by the nitrogen substitution. The sodium and cesium ions as "hard acids", cannot interact as strongly with the substituted nitrogen atoms of the ligand as they can with the oxygen atoms. The effect of the nitrogen substitution on the lithium and thallium(I) ion complexes is exactly the opposite, the stabilities of these complexes are greatly increased. These cations can form partially co- valent bonds which cause an increase in the strength of the interaction between the ligand and the cations upon the nitrogen substitution. Mojtaba Shamsipur The kinetics of the complexation reactions of the cesium ion with DB2lC7, DB2MC8, and DBBOClO in acetone and methanol were investigated by the temperature dependent 133Cs NMR. The energies of activation for the release of Cs+ from the CESiUfll complexes decrease with decreasing donicity of the solvent as expressed by the Gutmann donor number. They also decrease with the increase in the size of the ligand. The data show that, first, the transition state must involve a substantial ionic solvation and, second, the transition state must be more ordered than the initial and the final states, iii; the solvated complex and the solvated cesium ion and the free ligand. To Nahid ii I . ACKNOWLEDGMENTS The author wishes to express his sincere gratitude to Professor Alexander I. Popov for his guidance, encourage- ment, and friendship throughout this study. Professor Stanley R. Crouch is acknowledged for his many helpful suggestions as second reader. The author acknowledges the financial assistance of the people of Iran, as administered by the Isfahan Uni- versity of Technology, during the course of this study. The financial aids of the Department of Chemistry, Michigan State University, and the National Science Foundation are also acknowledged. The help of Mr. Frank Bennis, Mr. Wayne Burkhardt, and Mr. Tom Clarke in keeping the NMR spectrometers in operat— ing condition is acknowledged. Deep appreciation to my wife, Nahid, for her love, understanding, patience, and encouragement throughout this study. To her and to our son, Ali, I dedicate this thesis. iii ‘9 ‘ hh“ ._, Vt‘n‘ ‘ TABLE OF CONTENTS Chapter Page LIST OF TABLES. . . . . . . . . . . . . . . . . . .viii LIST OF FIGURES . . . . . . . . . . . . . . . . . . xv LIST OF ABBREVIATIONS . . . . . . . . . . . . . . .xxii CHAPTER 1. HISTORICAL REVIEW . . . . . . . . . . . 1 1.1. Macrocyclic Crown Ethers. . . . . . . . . 2 1.1.1. Introduction. . . . . . . . . . . 2 1.1.2. Metal Ion Complexes with Large Crowns. . . . . . . . . . . 3 1.1.3. Thermodynamics of Metal— Complex Formation in Solution . . 1“ 1.1.3.1. Open Chain Ligands. . . 1h 1.1.3.2. Cyclic Ligands. . . . . 21 1.2. Nuclear Magnetic Resonance. . . . . . . . 27 1.2.1. Introduction. . . . . . . . . . . 27 1.2.2. Chemical Shift Measurements . . . 28 1.2.3. Multinuclear NMR Studies of Complexation of T1+ and Alkali Ions in Solution. . . . . . . . 33 1.3. Conclusions . . . . . . . . . . . . . . . MO CHAPTER 2. EXPERIMENTAL PART . . . . . . . . . . . ill 2.1“. Synthesis and Purification of Ligands . . . . . . . . . . . . . . . . . 42 2.1.1. Synthesis of Dibenzo-30- crown-100 o o o o o o o o o o o o “2 2.1.2. Purification of Ligands . . . . . H3 2.2. Solvents and Salts. . . . . . . . . . . . All 2.2.1. Solvents. . . . . . . . . . . . . A” iv Chapter 2.3. 2.14. 2.5. CHAPTER 3. 3.3. 3.“. 3.5. Page 2.2.2. Salts . . . . . . . . . . . . . . 45 Sample Preparation. . . . . . . . . . . . M6 Instrumental Measurements . . . . . . . . 46 Data Handling . . . . . . . . . . . . . . A8 MULTINUCLEAR NMR STUDY OF DIBENZO-30-CROWN-10, DIBENzo-2u-CROWN-8, AND DIBENZO—21-CROWN—7 COMPLEXES WITH Na+, K+, Cs+, and Tl+ IONS IN NONAQUEOUS SOLVENTS . . . . . . 50 Introduction. . . . . . . . . . . . . . . 51 Complexation of Na+, K+, and Cs+ Ions with Dibenzo-30-Crown-10 . . . . 52 3.2.1. DB3OC10 Complexes with Na+ and K+. . . . . . . . . . . . 52 3.2.2. DBBOCIO Complexes with Cs+. . . . 59 Complexation of Na+, Cs+, and T1+ Ions with Dibenzo-24-Crown—8. . . . . 67 3.3.1. DB2uc8 Complexes with Na+ . . . . 67 3.3.2. DB2uc8 Complexes with Cs+ . . . . 72 3.3.3. DB2uc8 Complexes With T1+ . . . . 77 3.3.u. Conclusions . . . . . . . . . . . 83 + + + Complexation of Na , Cs , and T1 Ions with Dibenzo-2l-Crown-7. . . . . . . 86 3.u.l. DB2107 Complexes with Na+ . . . . 86 3.u.2. DB21C7 Complexes with Cs+ . . . . 91 3.u.3. DB2107 Complexes with T1+ . . . . 97 3.u.u. Conclusions . . . . . . . . . . . 102 Discussion. . . . . . . . . . . . . . . . 105 Chapter Page CHAPTER A. CESIUM-133 NMR STUDY OF THE THERMODYNAMICS OF THE COMPLEXA- TION OF DIBENZO-30-CROWN—10, DIBENZO-2H—CROWN-8, AND DIBENZO— 21-CROWN—7 WITH CESIUM ION IN NONAQUEOUS SOLVENTS . . . . . . . . . . 109 4.1. Introduction. . . . . . . . . . . . . . . 110 u.2. DB3OC10 Complexes with 03+. . . . . . . . 111 u.3 DE2uc8 Complexes with Cs+ . . . . . . . . 126 u.u. DB2IC7 Complexes with Cs+ . . . . . . . . 139 u.5. Discussion. . . . . . . . . . . . . . . . 1&9 CHAPTER 5. LITHIUM-7, SODIUM-23, CESIUM-133, AND THALLIUM—205 NMR STUDY OF Li+, Na+, Cs+, and Tl+ ION COMPLEXES WITH 1,10-DIAZA-l8-CROWN-6 IN VARIOUS NONAQUEOUS SOLVENTS . . . . . . 156 5.1. Introduction. . . . . . . . . . . . . . . 157 5.2. Results . . . . . . . . . . . . . . . . . 158 5.2.1. 1,10—Diaza-18-Crown-6 Complexes with Li+. . . . . . . . 158 5.2.2. 1,10-Diaza—18-Crown-6 Complexes with Na+. . . . . . . . 170 5.2.3. 1,10-Diaza-18-Crown-6 Complexes with 05+. . . . . . . . 173 5.2.U. 1,10-Diaza—l8—Crown-6 Complexes with T1+. . . . . . . . 176 5.3. Discussion. . . . . . . . . . . . . . . . 177 CHAPTER 6. A.STUDY OF DYNAMICS OF CESIUM ION COMPLEXES WITH DIBENZO-30- CROWN-10, DIBENZO-2A-CROWN-8, AND DIBENZO-Zl-CROWN-7 IN ACETONE AND METHANOL. . . . . . . . . . 183 6.1 Introduction . . . . . . . . . . . . . . . 18A 6 (5.2. Determination and Interpretation of the Lineshapes . . . . . . . . . . . . 186 vi Chapter Page 6.3. Results and Discussion. . . . . . . . . . 195 APPENDICES APPENDIX I - DETERMINATION OF COMPLEX FORMATION CONSTANTS BY THE NMR TECHNIQUE, DESCRIPTION OF COMPUTER PROGRAM KINFIT AND SUBROUTINE EQUATION . . . . . . . . . . . . . . 211 APPENDIX II - DETERMINATION OF COMPLEX FORMATION CONSTANT WITH ION PAIR FORMA— TION BY THE NMR METHOD. . . . . . . . . . . . . 216 REFERENCES. . . . . . g . . . . . . . . . . . . . . 223 vii Table LIST OF TABLES Experimental Conditions for Metal Ion NMR . . . . . . . Thermodynamic Parameters for the Complexation of Large Crown Ethers with Cations in Solution. . . . Thermodynamic Data for Diethylene- triamine Complexes with Some Metal (II) Ions in 0.1 M KCl at 25°C. Thermodynamic Data for Metal (II)- Polyamine Complexes (Chelate Effect) . . . . . . . Thermodynamic Data for Reaction (2) for Different Metal (II) Ions Nuclear Properties of Alkali Elements and Thallium . . . . . . . Key Solvent Properties and Cor- rection for Magnetic Susceptibility on DA-60. . . . . . . . . . . . . Mole Ratio Study of Dibenzo-30- Crown—10 Complexes with 0.05 M, Sodium Tetraphenylborate in Various Solvents at 30°C. . . viii Page “7 11 13 l7 19 35 “9 53 Table 10 ll 12 13 Page Mole Ratio Study of Dibenzo-30— Crown-10 Complexes with 0.005 M Cs+ Ion in Various Solvents at 30°C. . . . . . . . . . . . . . . . . . . . 60 Formation Constants and the Limiting Chemical Shifts of DB3OC10-Cs+ Complexes in Various Solvents. . . . . . . . . . . . . . . . . . 66 Mole Ratio Study of Dibenzo-2h- Crown-8 Complexes with 0.025111 Sodium Tetraphenylborate in Various Solvents at 30°C. . . . . . . . . . 68 Formation Constants and the Limiting Chemical Shifts of DBZACB-Na+ Complexes in Various Solvents. . . . . . . . . . . . . . 71 Mole Ratio Study of Dibenzo-ZM- Crown-8 Complexes with 0.005 M Cs+ Ion in Various Solvents at 30°C . . . . . . . . . . . . . . . . . . 7“ Formation Constants and the Limiting Chemical Shifts of + DB2AC8-Cs Complexes in ‘Various Solvents. . . . . . . . . . . . . . 76 Table Page 1” Mole Ratio Study of Dibenzo-2H- Crown-8 Complexes with 0.005 M T1C10u in Various Solvents at 30°C. . . . . . . . . . . . . . . . . . . . 79 15 Formation Constants and the Limiting Chemical Shifts of DBZMCB-Tl+ Complexes in Various Solvents. . . . . . . . . . . . . . . . . . 82 16 Formation Constants of 1:1 Com- plexes of Na+, Cs+, and T1+ Ions with Dibenzo-Zu-Crown-8 in Various Solvents . . . . . . . . . . . . 8H 17 Mole Ratio Study of Dibenzo-21— Crown-7 Complexes with 0.025 M NaBPhu in Various Solvents at 30°C. . . . . . . . . . . . . . . . . . . . 87 18 Formation Constants and the Limiting Chemical Shifts of DB21C7 -Na+ Complexes in Various Solvents. . . . . 92 19 Mole Ratio Study of DibenzO-21- Crown-7 Complexes with 0.005 M CsSCN in Various Solvents at 30°C. . . . . . . . . . . . . . . . . . . . 9h 20 .Formation Constants and the Limiting Chemical Shifts of Table Page DB21C7-Cs+ Complexes in Various Solvents. . . . . . . . . . . . . . . . . . 96 21 Mole Ratio Study of Dibenzo-21- Crown-7 Complexes with 0.005 M T1C10u in Various Solvents at 30°C. . . . . . . . . . . . . . . . . . . . 98 22 Formation Constants and the Limiting Chemical Shifts of DB21C7-T1+ Complexes in Various Solvents. . . . . . . . . . . . . . 101 23 Formation Constants of 1:1 Com- plexes of Na+, 08+, and T1+ Ions with DB21C7 in Various Solvents . . . . . . 103 24 Formation Constants of 1:1 Complexes of Na+, 05+, and T1+ Ions with DB21C7, DBZ4C8, and DB30010 in Various Solvents at 30°C . . . . . . . . 106 25 Cesiump133 Chemical Shifts of 0.005 M Cs+ Ion in the Presence of DB30010 at Various Temperatures. . . . . . . . . . . . . . . . 113 25 Formation Constants of DBBOCIO 'Cs+ Complex in Nonaqueous Sol— vents at Different Temperatures . . . . . . 122 xi Table 27 28 29 30 31 32 33 Thermodynamic Parameters for the Complexation Of Cs+ Ion by Dibenzo- 30-Crown-10 in Various Solvents Cesium-133 Chemical Shifts of 0,005 M.CS+ Ion in the Presence of DB24C8 at Various Temperatures Formation Constants of DB24C8-Cs+ Complex in Nonaqueous Solvents at Different Temperatures. . . . . . Thermodynamic Parameters for the Complexation of Cs+ Ion by Dibenzo-24-Crown-8 in Various Solvents. . . . . . . . . . . . . Cesium-133 Chemical Shifts of 0.005 Mos+ Ion in the Presence of DB21C7 at Various Temperatures Formation Constants of DBZlC7-Cs+ Complex in Nonaqueous Solvents at Various Temperatures. . . . . . Thermodynamic Parameters for the Complexation Of Cs+ Ion by DBZlC? in Nonaqueous Solvents at Different Temperatures . . . . xii Page 124 127 136 138 140 148 151 Table 34 35 36 37 38 39 40 Entropies of the Complexation of Cesium Ion by DB3OC10, DB24C8, and DB21C7 in Various Solvents. . . . . . . . Mole Ratio Study of 1,10-Diaza-18- Crown-6 Complex With 0.02 M LiClOu in Various Solvents at 30°C . . . . . . . Mole Ratio Study of 1,10-Diaza— 18-Crown-6 Complex with 0.05 M NaBPhu in Various Solvents at 30°C. . . . . . . . . . . . . . . . . . . Mole Ratio Study of 1,10-Diaza- 18-Crown-6 Complex with Cs+ Ion in Various Solvents at 30°C . . . . . . . Mole Ratio Study of 1,10-Diaza-18- Crown-6 Complexes with T1+ Ion in Various Solvents at 30°C . . . . Formation Constants and the Limiting Chemical Shifts of 1,10-Diaza-18-Crown-6-Li+ Complexes in Various Solvents Formation Constants and the Limiting Chemical Shifts Of 1,10-Diaza-18-Crown—6-Na+ Complexes in Various Solvents . . xiii Page 153 159 160 161 163 169 172 Table 41 42 43 44 45 Page Formation Constants and the Limiting Chemical Shifts of 1,10-Diaza-18-Crown-6-Cs+ Complexes in Various Solvents . . . . . . . 175 Formation Constants and the Limiting Chemical Shifts of 1,10-Diaza-18-Crown-6'T1+ Complexes in Various Solvents . . . . . . . .178 Formation Constants of Li+, Cs+, and T1+ Ion Complexes of 1,10- Diaza—lB—Crown-6 and 18-Crown-6 in Various Solvents . . . . . . . . . . . . 179 Temperature Dependence of the Rate Constants for the Release of Cs+ Ion from D821C7-Cs+, DB24C8-Cs+, and DB3OC10-Cs+ Complexes in Acetone and Methanol. . . . . . . . . . . . 202 Exchange Rates and Thermodynamic Parameters for Release of Cs+ Ion from Some Large Crown Complexes in Acetone and Methanol. . . . . . . . . . . . . . . . . . 206 xiv Figure LIST OF FIGURES Page Structure of Some Large Crown Ethers. . . . . . . . . . . . . . . . . . 4 Crystalline Structure of Some Metal Ion-Crown Complexes. A—(KSCN)2-DB2408, B—KI-DB30C10, C—(NaSCN)2DB30C10 . . . . . 7 Sodium-23 Chemical Shifts YE: [DB30C10]/ [Na+1 Mole Ratio in Different Solvents. A-Nitromethane, B-Acetonitrile, C- Pyridine. . . . . . . . . . . . . . . . . 54 Carbon-13 Chemical Shifts at Various [Metal IonJ/[DB30C10] Mole Ratios (MR) in Nitromethane. A-Sodium Ion at MR=0.0, 1.0, and 2.0; B- Potassium Ion at MR=0.0, 0.5, and 1.0 . . . . . . . . . . . . . . . . . . . 56 Carbon-13 Chemical Shifts for the Four Polyether Chain Carbon Atoms at Various [Na+1/[DB3OC10] Mole Ratios in Nitromethane . . . . . . . 57 Cesium-133 Chemical Shifts 1s [DB3OC10]/[Cs+] Mole Ratio in XV Figure 10 ll 12 13 Page Different Solvents. A—Nitromethane, B-Methanol, C-Acetone, D-Pyridine, E—Acetonitrile. . . . . . . . . . . . . . 61 Computer Fit of the Cesium-133 Mole Ratio Data for DB3OC10-Cs+ in Methanol at 30°C. . . . . . . . . . . . . 65 Sodium—23 Chemical Shifts Mg. [DB24C8]/[Na+] Mole Ratio in Different Solvents. . . . . . . . . . . . 70 Cesium—133 Chemical Shifts Kg, [DBZ4C8j/[Cs+] Mole Ratio in Different Solvents. . . . . . . . . . . . 75 Thallium-205 Chemical Shifts Kg. [DB21C81/[T1+] Mole Ratio in Dif- ferent Solvents . . . . . . . . . . . . . 80 Sodium-23 Chemical Shifts gs. [DB21C7]/[Na+] Mole Ratio in Different Solvents. . . . . . . . . . . . 89 Cesium-133 Chemical Shifts Kg. [DB21C7]/[Cs+] Mole Ratio in Different Solvents. . . . . . . . . . . . 95 Thallium-205 Chemical Shifts gs. [DBZlC7]/[T1+] Mole Ratio in Different Solvents . . . . . . . . . . 99 xvi Figure 14 15 16 17 18 19 2O 21 Page Cesium-133 Chemical Shifts 1g. [DB3OC10]/[Cs+] Mole Ratio in Nitromethane at Different Tem- peratures . . . . . . . . . . . . . . . . 116 Cesium-133 Chemical Shifts Kg. [DB3OC10]/[Cs+] Mole Ratio in Acetonitrile at Different Tem- peratures . . . . . . . . . . . . . . . . 117 Cesium-133 Chemical Shifts YE- [DB3OClO]/[Cs+] Mole Ratio in Acetone at Different Temperatures . . . . 118 Cesium-133 Chemical Shifts gs. [DB3OClO]/[Cs+] Mole Ratio in Methanolat Different Temperatures . . . . 119 Cesium-133 Chemical Shifts gs. [DB3OClO]/[Cs+] Mole Ratio in Pyridine at Different Temperatures. . . . 120 Van't Hoff Plots for Complexation of Cs+ Ion by Dibenzo-30-Crown-10 in Various Solvents. . . . . . . . . . . . . 123 Cesium-133 Chemical Shifts lg. [DB2408]/[Cs+] Mole Ratio in Nitro- methane at Different Temperatures . . . . 130 Cesium-133 Chemical Shifts 1g. [DB24C8]/[Cs+] Mole Ratio in xvii Figure Page Acetonitrile at Different Tem- peratures . . . . . . . . . . . . . . . . 131 22 Cesium-133 Chemical Shifts gs. [DB24C8]/[Cs+] Mole Ratio in Acetone at Different Temperatures . . . . 132 23 Cesium—133 Chemical Shifts lg. [0324081/[Csfj Mole Ratio in Methanol at Different Temperatures. . . . 133 24 Cesium-133 Chemical Shifts Kg. [DBZ4C8]/[Cs+] Mole Ratio in Pyridine at Different Temperatures. . . . 134 25 Van't Hoff Plots for Complexation of Cs+ Ion by Dibenzo-24—Crown-8 in Various Solvents . . . . . . . . . . . 137 26 Cesium-133 Chemical Shifts vs. [DB21C7]/[Cs+] Mole Ratio in Nitromethane at Different Tem- peratures . . . . . . . . . . . . . . . . 143 27 Cesium-133 Chemical Shifts YE- [DB2107]/[Cs+] Mole Ratio in Acetonitrile at Different Tem- peratures . . . . . . . . . . . . . . . . 144 28 Cesium-133 Chemical Shifts YE- [DBZlC71/[Cs+] Mole Ratio in Acetone at Different Temperatures . . . . 145 xviii Figure Page 29 Cesium-133 Chemical Shifts 1s. [DB21C7]/[Cs+] Mole Ratio in Methanol at Different Temperatures. . . . 146 30 Cesium-133 Chemical Shifts 1g. [DBZlC7]/[Cs+] Mole Ratio in Pyridine at Different Temperatures. . . . 147 31 Van't Hoff Plots for Complexation of Cs+ Ion by Dibenzo-21-Crown-7 in Various Solvents . . . . . . . . . . . 150 32 Lithium-7 Chemical Shifts vs. [DA18C6J/[Li+] Mole Ratio in Different Solvents. . . . . . . . . . . . 16“ 33 Sodium-23 Chemical Shifts gs. [DA18C6J/[Na+] Mole Ratio in Different Solvents. . . . . . . . . . . . 165 34 Cesium-133 Chemical Shifts YE- [DA18C6J/[Cs+] Mole Ratio in Different Solvents. . . . ._. . . . . . . 166 35 Thallium-205 Chemical Shifts gs. [DA18C61/[T1+] Mole Ratio in Different Solvents. . . . . . . . . . . . 167 36 Cesium-133 NMR Spectra of 0.02 M CsSCN, 0.01 M_DBZlC7 Solution in Acetone at Various Temperatures . . . . . 196 xix Figure Page 37 Cesium-133 NMR Spectra of 0.04 M CsSCN, 0.02 M DB2107 Solution in Methanol at Various Temperatures. . . . . 197 38 Cesium-133 NMR Spectra of 0.02 M CsSCN, 0.01 M DB24C8 Solution in Acetone at Various Temperatures . . . . . 198 39 Cesium-133 NMR Spectra of 0.02 M_ CsSCN, 0.01 M DB2408 Solution in Methanol at Various Temperatures. . . . . 199 40 Cesium-133 NMR Spectra of 0.01 M CsSCN, 0.005 M_DB30C10 Solution in Acetone at Various Temperatures. . . . 200 41 Cesium-133 NMR Spectra of 0.01 M CsSCN, 0.005 M DB3OC10 Solution in Methanol at Various Temperatures . . . 201 42 Arrhenius plots of log kb 172. l/T for the Release of Cs+ Ion in Acetone and Methanol with Large Crown Ethers. A-DB30C10-Cs+ in Methanol, B-DB2408-Cs+ in Methanol, C-DB3OClo-Cs+ in Acetone, D-DB21C7'Cs+ in Meth- anol, E-DBZ4C8'Cs+ in Acetone, F-DB21C7°CS+ in Acetone . . . . . . . . . 206 XX Figure Page 43 Entropy Profile for the Release + 'Of Cs Ion from DB3OC10-Cs+ Complex in Acetone. . . . . . . . . . . . . . . . 209 xxi Ac AN DMF DMSO MeOH NM PC Py TMG TMO DB3OC10 DB24C8 DB2107 DA18C6 LIST OF ABBREVIATIONS Acetone Acetonitrile Dimethylformamide Dimethylsulfoxide Methanol Nitromethane Propylene carbonate Pyridine Tetramethylguanidine Trimethylene oxide Dibenzo-30-crown-10 Dibenzo-24-crown-8 Dibenzo-21-crown-7 1,10-Diaza-l8-crown-6 xxii CHAPTER 1 HISTORICAL REVIEW 1.1. MACROCYCLIC CROWN ETHERS 1.1.1. Introduction Since Pedersen's discovery of macrocyclic polyether (crown) compounds capable of forming stable complexes with the alkali ions (1) the studies of these ligands and their complexes have become a very popular field of research. A large number of such complexes have been isolated in crystalline form and many solution studies have been car- ried out (2,3). Most of the investigations in solution have been done in water, methanol and/or their mixtures; such studies in other nonaqueous solutions are quite sparse. A variety of physicochemical techniques have been used for such investigations (3); the choice of favorite technique being dictated by the systems studied as well as by the particular expertise of the investigators. One of the most interesting characteristics of the macrocyclic compounds is their ability to selectively bind certain cations in solution in the presence of others. The selectivity and stability of crown ether complexes has been reported to be dependent on several important parameters characteristic of the ligand, the cation and the reaction medium. These parameters are: the relative sizes Of cation and ligand cavity (4,5), the type and the number of donor atoms in the ring (4,6,7,8), substi- tution on the macrocyclic ring (9-12), type and charge 2 of cation (13) and the solvent effect (14-16). Several useful review articles are available on the study of macrocyclic polyethers and their complexes (2,3, 17-19). In this thesis only the studies on the complexes of large crowns and the thermodynamics Of the metal- complex formation in solution will be reviewed. 1.1.2. Metal Ion Complexation with Large Crowns Despite the very interesting properties of large crowns (143;) larger than 18-crown-6), not much attention has been focused on the study of the complexation of metal ions with these ligands. In comparison with approximately three hundred scientific papers on small size crowns, and in particular on different 18—crown-6 derivatives, such investigations on large crown complexes are sparse. The structure of some of these ligands are shown in Figure 1. Pedersen (1) was the first to report the isolation of cesium ion complexes with dibenzo-24-crown-8 and dibenzo- 30-crown-10 in methanol solution and to study the effect of complexation on the ultraviolet spectra of large crowns. He likewise (20) studied complexation of alkali and alkali earth cations with 21-crown—7 and 24-crown-8 derivatives using ultraviolet spectroscopy and solvent extraction methods. The effect of dicyclohexy1-24-crown—8 on the ionic permeability of natural membranes such as HK and LK sheep DIBENZO—Ql-CROWN—7 Figure 1. Structure of Some Large Crown Ethers. red cells has been studied by Tosteson (21). He showed that this ligand notably increases potassium permeability more than sodium permeability, while the behavior is opposite in the case of cyclohexyl-15-crown-5. It has been reported (22) that cyclic polyethers exhibit specific influences on cation transport in rat liver mitochondria. Dicyclohexy1—2l-crown-7 was most effective in the presence of potassium or rubidium ions. Dicycloxyl-BO-crown-lo was displayed striking specifity in that it was very active with rubidium present and much less active with potassium. The stoichiometry of crystalline complexes of alkali cations with large crowns has been investigated (23-25), and the crystalline structure of some of these complexes has been reported (26—29). The stoichiometry of such compounds depends on the relative sizes of the cation and the cavity size of the macrocycle, the flexibility of the crown molecule, and the nature of the anion and of the solvent (17). For dibenzo-24-crown-8—K+ and dibenzo-30- crown-lo—Na+ systems the ratio of two metal ions to one crown ether has been found so far (24). Mercer and Truter (27) have reported the structure of the 2:1 (metal to ligand) complex between KSCN salt and dibenzo-24-crown-8, isolated from methanol solution. According to them, each one of the K+ ions, located in— side the ring, are bound to the five oxygen atoms of the ligand, two nitrogen atoms of bridging thiocyanate groups, and two carbon atoms of a benzene ring (Figure 2a). In another publication (25). Truter and co-workers have shown that when potassium tetraphenylborate was used instead Of potassium thiocyanate, only the 1:1 complex was formed since the tetraphenylborate anion cannot be coordinated with the metal ion. The crystalline structure of the 1:1 complex between KI and dibenzo-30-crown—10 has been studied by Bush and Truter (26). They showed that the potassium ion is com- pletely located inside the cavity which is created by the twisting of the ligand around the cation, all the oxygen atoms are coordinated with the cation to form a so called "wrap around" complex (Figure 2b). Owen and Truter (29) recently have reported the crystal— line structure of the 2:1 (metal to ligand) complex of sodium isothiocyanate with dibenzo-30-crown-1 (29), determined from x-ray diffraction measurements. Each ligand complex with two Na+ ions, and each cation is co- ordinated to six oxygen atoms in the ring and also to one isothiocyanate anion, through the nitrogen atom. The ligand is in an extended conformation, twisted so that the molecule assumes a figure eight configuration, with the Na+ ions at the center of two loops (Figure 2c). The stability constants for the 1:1 complexes of large crown ethers (i.e., 21, 24, 30, and 60-membered rings) ‘73" ’fih 1W? x ‘ a 1 ' @53c 75);} U“... ‘ C Figure 2. Crystalline Structure of Some Metal Ion-Crown Complexes. A-(KSCN)2-DB24C8, B-KI-DBBOClO, C-(NaSCN)2DB30010. with alkali ions has been determined by Frensdorff (4) in water and methanol solutions using potentiometric measurements with cation-selective electrodes. He point- ed Out that the selectivity order is governed by the rela- tive sizes of the cation and the cavity of the ligand. From the increase of K+ stability constants between 24- crown-8 and 30-crown-10 it was suggested that a "wrap around" complexing might occur with polyethers, such as that observed for K+-valinomycin (30). Rechnitz and Eyal (31) constructed liquid membrane electrodes with dibenzo—30-crown-10 in the membrane phase (nitrobenzene) to measure the crown's electrochemical activity, and to determine the crown-metal ion complex formation constant. The potentiometric selectivity ratio for the crown ether in nitrobenzene for the Rb+, 03+, Na+, and NH”+ ions with respect to K+ ion has been de- termined, and formation constants of rubidium and potassium ion complexes with dibenzo-30-crown-1O have been reported. The AG° and AH° parameters for complex formation of di- benzo-30-crown-10 and Li+, Na+, K+, Rb+ , Cs+, NHu+ and Tl+ in methanol has been determined by Chock (32), who has studied the ultraviolet spectra of the ligand at different temperatures. With the exception of the am- monium ion, a good correlation between the stability constant of the complex and the size of metal ion was observed. In an attempt to elucidate the solution structure Of a number of crown ethers, Live and Chan (33) carried out a careful measurement on the 1H and 13C NMR spectra of the ligands and their metal complexes in water, water- acetone, acetone, and chloroform solutions. From the data, the authors have concluded that the structure of K+ dibenzo-30-crown-10 in solution is similar to that in crystalline state (26) where K+ ion is completely surrounded by the ligand to form a "wrap around" complex. 0n the basis of the proton NMR spectrum of Na+ dibenzo-30-crown- 10 complex in acetone, which did not resemble the one for + K , they proposed a twisted configuration for the complex in solution. Their 13C NMR results support these conclu- sions. The stability constants for 1:1 complexes of the + ligand with Na+, K , and Cs+ in acetone have been deter- 1 mined from the H NMR data. Izatt and co-workers (34) have reported thermodynamic +, Rb+, and Cs+ ions parameters for complexation Of Na+, K with dibenzo-24-crown-8 and dibenzo-27-crown-9 in methanol- water mixtures which have determined calorimetrically. They pointed out that the entropy changes for a 1:1 re- action of a given cation become more negative with in- creased size of the polyether ring which seem to indicate an increased conformational change of the ligand upon complex formation. The stability of thallium and alkali metal ion C) .9 w- :i I‘ir 10 complexes with dibenzo-24-crown-8, dibenzo-30—crown-10 and their benzo group derivatives in acetonitrite and methanol has been investigated by Mittal and co-workers using polarographic methods (35). The nature of the sol- vent was found to affect the stability constants of the complexes. Recently, Lehn and co-workers (36) reported formation of stable and selective complexes between guadinium and imidazolium ions and chiral hexacarboxylate-27-crown-9 in aqueous solution. The stability constants of the complexes were measured potentiometrically by a competition method using a NHu+-selective electrode. They have con- cluded that the presence of anionic carboxylate groups in the ligand markedly increases the stability of the crown complexes and that the selectivity of the complexa- tion arises from central discrimination of the macrocyclic ring. The literature reported thermodynamic values for the complexation reaction of metal ions with large crowns are listed in Table 1. ((AI\ fir .* CL A i W.- ruSIKrA .rle {Adv/KPFC CFHAF ~\:H LC :OwecxoacEco ecu Lo; nLOBOEmkmL OHEmEAUOELO£E .H OHQTR 11 mm IIII IIII om.m woanoa z< +mo :m m.~I m>.mI mm.m and mom: Ron +nm am m.~H :m.mI m:.m ado mom: mow : IIII IIII m:.m pod moo: mm IIII IIII om.m wonmaoa mom: mm IIII IIII os.m mohoaoa za +x am m.mH m~.~I :m.a and mom: go» mm IIII IIII 00.: mommaoa z< +mz monmma : IIII IIII m.H pom nouns +mo mozmoa : IIII IIII ma.= poo mom: +mo a IIII IIII m:.m pom mom: +m mozm z IIII IIII om.= poo mom: +mo : IIII IIII om.= pod moo: +x : IIII IIII :.m pom mom: +mz soammn : IIII IIII m.H pom noun; +mo Noamon : IIII IIII mo.m pom moms +mo : IIII IIII H=.: poo mom: +x Noam .mmm Ammo mHoE\Hwov AmHoE\Hmoxv m woq vogue: pcm>aom coapmo pcmwaq m< m< .COHpsaom CH mCOHpmo spas mnmcpm czono mmnmq no coaummequo on» pom mnmumemnmm Odemcmpoenmne .H manna l2 mm =.HNI s.mHI no.2 coon moo: +om Hm IIII IIII mm.H sod axe Rom mm IIII IIII om.s m:z oa mm IIII IIII os.= moaoaoo 2: mm IIII IIII om.m woaoaoa :oo: mm s.sHI m.HHI sm.s soon :oo: o IIII IIII oo.: poo moo: +: mm IIII IIII sm.m m:z o< mm IIII IIII om.m wopoaoq 2: mm s.mI o.aI HH.N coon moo: a IIII IIII o.m cod moo: +oz oaoommo am H.aHI sa.oI ma.H Hoe moo: sos +no am m.mHI om.mI om.m Hoe moo: mos +: am m.mmu es.HHI om.H Hoe moo: nos +oz mosmoo mm IIII IIII o=.m wosoaoo :oo: mm IIII IIII om.a mosoaoa z< +He am o.mHI mm.mu m=.m Hoe moo: nos 3 IIII IIII ms.m pod moo: +no .mom Ammo oHoE\HMoV AoHoE\Hmoxv mm wog cospoz pzo>aom cowuwo ocwwfiq me me .ooasapcoo .H oHooe -EQZCfiuCCU . H THENCE ill] 13 a IIII IIII om.m sod moo: +: oaooooo mm m.sI m.mI me.m ooan moo: +e:z mm IIII IIII oo.m moaofioa z< mm IIII IIII oa.a wotofloa :oo: mm m.oHI o.HHI Hm.= coda moo: +He mm IIII IIII mm.s m:z o< mm IIII IIII om.m wosofioo ze mm «.mHI N.HHI mm.= ooao moo: +no Hm IIII IIII om.H cod axe mom mm IIII IIII os.s woooaod z< +om .rumm Awmv mHOE\HmoV AOHOE\H.®0MV MM wOA UOSQOE me>Hom £0.“me Ucmwfiq no mo .ooscaocoo .H oases Wei] 6 E. r x n... 63’ It We A Ir .. LIV 14 1.1.3. Thermodynamics of Metal-complex Formation in Solution 1.1.3.1. Open Chain Ligands - The thermodynamics of metal ion complexes of a number of polyamine systems have been investigated calorimetrically by Paoletti g£_a1. (37-40) in aqueous solutions. The ligand studies were diethylene triamine (dien) (30), n=l; triethylene tetra- mine (trien) (38), n=2; tetraethylenepentamine (tetren) (39), n=3; and N,N,N',N'-tetra-(2-aminoethyl)-ethylene- diamine (Penten) (40). Some of the thermodynamic data for these complexes are listed in Table 2. H 2 Polyamines' Structure It is seen that there is a tendency for successive enthalpy changes to become more exothermic, while the entropy changes for 1:1 complexes are positive, those for the second step reactions are negative. The overall entropy changes, however, are positive in all cases and the complexes are both enthalpy and entropy stabilized. The nature of the chelate effect can be examined in more detail using the available thermodynamic data for the polyamine complexes for the following reaction. 15 Table 2. Thermodynamic Data for Diethylenetriamine Com- plexes with Some Metal (II) Ions in 0.1 M KCl at 25°C. 2+ AG AH AS M (Kcal/mole) (Kcal/mole) (cal/mole deg) Co2++dien -10.90 —8.15 9.0 Codien2++dien -8.00 -10.25 -7.5 Niz++dien -14.45 -11.85 8.5 Nidien2++dien -10.90 -13.45 —8.5 Cu2++dien -21.55 -l8.00 12.0 Cudien2++dien -7.10 -8.15 -3.5 Zn2++dien -12.00 -6.45 18.5 anien2++dien -7.50 -10.15 -9.0 i. I; :x AV 16 2+ M(NI-IBM + L z ML2+ + n NH3 (1) Some of the results obtained are shown in Table 3. It can be seen that the complexes are stabilized with respect to the corresponding ammonia complexes not only by enthalpy changes but also by the large positive en- tropy changes for the reactions. From the data, it is clear that the large negative free energy changes can be attributed to the positive entropy changes accompanying the release of increasing numbers of ammonia molecules. Extensive calorimetric studies have been made Of the formation of metal complexes with aminopolycarboxalate ions such as ethylendiamine tetraacetic acid and its homolagues (42-48). Schwarzenbach (41) has shown that in ligands with more than two donor atoms the presence of a nitrogen atom meets the requirements for the forma- tion of relatively strain-free chelate rings. These ligands have been found to be stabilized by positive entropy changes resulting from charge neutralization and subsequent solvent release from the solvation shells of interacting ions. Thermodynamics data are obtained from those available for methyliminodiacetate (mIDA) (42) and ethylendiamine tetraacetate homologues: 17 Table 3. Thermodynamic Data for Metal (II)-Polyamine Complexes (Chelate Effect). AS 2+ AG AH (cal/mole) M L (Kcal/mole) (Kcal/mole) deg) Ref. 2+ Ni dien -5.33 -1.35 13.4 37 Cu2+ dien -7.24 -3.00 14.2 37 Ni2+ trien -8.09 0 27.1 38 Cu2+ trien -lO.20 -l.55 29.0 38 N12+ tetren -12.10 -1.40 35.9 39 Cu2+ tetren -14.65 —1.75 43.3 39 Ni2+ penten -15.44 +1.35 56.0 40 o. .. . in «a I? v . w r .. I... o .I. I I C to F. In. l8 OOCCH2 l/CH2COO _ N+ 2mc2 O K r3 O K 2(E -E )'1[< lzi X |z 2K] >1 0 m. ll’0 me wm ‘3 YO m K K rK where e : electronic charge n1: mass of the electron c : velocity of light EK : angular momentum of the Kth electron th electron from the Kth rK : radial distance Of the K electron from the origin at the nucleus. The diamagnetic component, Gd, depends on ground state electronic wave functions and is a function of the sym- metry of the electronic distribution and the density of circulating electrons. The magnitude of the paramagnetic contribution, up, is zero for ions with spherically sym— metrical S states but it is substantial for atoms involved in chemical bonding. It is determined by several factors. (1) The inverse of the energy separations 1E between ground and excited electronic states of the molecule. (ii) The relative electron densities in the various p, d, and higher states involved in bonding, i;g;, upon the degree of asymmetry in electron distribution near the nucleus. 31 (iii) The value of <1/r3>, the average inverse cube dis- tance from the nucleus to the orbitals concerned. Usually the contributions to ad and CD for a nucleus are considered only for the electrons immediately neighboring, or local to, that nucleus. More distant electrons give rise to long range effects on both op and 9d which are large but cancel to make only a small net contribution to 0. Gen- erally, downfield shifts are referred to as paramagnetic and upfield shifts as diamagnetic. Kondo and Yamashita (73) have proposed the theory of paramagnetic interaction. They suggested that the para- magnetic shift of cations and anions in alkali halide crystals is due to the short range repulsive forces between the closed shell of the ions. These forces can excite p orbital electrons of the alkali nuclei to the higher states, so that the net result would be a decrease in the shield- ing of the nucleus. The success of the Kondo-Yamashita theory in inter- preting chemical shifts in solids suggested that it may also provide some way for interpretation of the chemical shifts in solution. In this case, however, the problem is more complex. In solids the relative positions and distances of separation of the ions are known, but in the solution the environment Of the nucleus will vary randomly with time because of the diffusion of the ions and solvent molecules through the solution and the observed chemical 32 shift will result from an average of many instantaneous values. Deverell and Richards (74) applied Kondo-Yamashita theory to provide a qualitative interpretation of the cation chemical shifts in aqueous solutions. They sug- gested that at infinite dilution, where the only inter- actions present are between the ion and water molecules, the contribution to the paramagnetic chemical shift is °=— 2_1_ 0l-0 an 16“ (r3>np A A ion-water where a is the fine-structure constant, ' 1. Since the <3;> and l-both increase with 1.3 np K r3 A increasing atomic number (75), the magnitude and the range 'I- of 0 increases from Li+ to Cs ions. Therefore, the range P of chemical shift varies from about 10 ppm for Li+ to several hundred ppm for Cs+ ion. 1.2.3. Multinuclear NMR Studies of the Complexation of T1+ and Alkali Ions in Solution During the past decade the use of multinuclear NMR for the studies of the thermodynamics and kinetics of reactions in solutions has been expanded very rapidly. In particular, multinuclear NMR has been widely used to study the behavior of alkali ions in solutions. Since the advent Of new macrocyclic ligands, such as crowns, discovered by Pedersen (l), and cryptands, dis- covered by Lehn (76), capable Of forming strong complexes with alkali ions, the studies of the alkali ion complexes with these ligands have become a very popular field of research. Among a variety of physicochemical techniques used for such investigations, alkali metal NMR has been shown to be very popular. 34 The nuclear properties of alkali elements as well as of thallium are shown in Table 5. All alkali nuclei have at least one isotope with non-zero spin. In all cases I is equal or greater than 1/2 and, therefore, the nuclei have a quadropole moment. Thus it should be expected that the alkali resonances would have broad lines. In prac- tice, however, the natural line width of 23Na and specially of 7Li, 133Cs and 205T1 nuclei are quite narrow so that in most cases chemical shifts can be measured precisely. Relative sensitivities are generally adequate for all of them except 39K. Thallium-205 is shown to be an ideally suited NMR probe for potassium in biological systems. The chemical properties and ionic radii (1.44 A and 1.54 A) Of K+ and Tl+ are similar so that T1+ can replace K+ in several enzymes without loss Of activity. Lithium-7 NMR has been used for determining formation constants of lithium complexes with pentamethylentetrazole in nitromethane (77). It was found that lithium ion forms a fairly strong complex with a convulsant tetrazole in nitromethane. Lithium ion complexes with cryptands C222, C221 and C211 in water and in several non-aqueous solvents have been studied by Cahen 92,3l. using 7Li NMR technique (78). They showed that the first two ligands form weak 1:1 com- plexes with Li+ ion in solvents of low donicity such as nitromethane. On the other hand, cryptand 211 was found 35 mmH.o N\H w=.os mam.:m Hamom NIOH x 35.: m\~ ooa 3mm.» .a..omm._H uwa.o m\m m.m~ omm.mH nmwm :IoH x mo.m _ m\m mo.mm oom.m Mam NIOH x pm.m m\m ooa www.ma ome :mm.o m\m nm.m: mam.mm HAN oaofim unopmcoo on ma :HQm ANV mmsomx msoaosz on o>HumHom mpa>aufimcom oocoocsn< mo.:H no Aumzv Annapmz mocoooonm mzz .ESHHHmnB oco mucoEon HmeH< mo mofiunoqonm nmoaosz .m manoe 36 to form much more stable complexes and two 7Li resonances (corresponding to the free and the complexed Li+) were observed for solutions containing excess of the Li+ ion. The resonance of the Li+ ion inside the cryptand cavity was found to be completely independent of the solvent in- dicating that the ligand completely insulates the cation from the solvent. The kinetics of the complexation reaction of the Li+ ion with cryptand C211 in water and several non-aqueous solvents have been investigated by temperature—dependent 7Li NMR (79). The activation energy for the release of Li+ from the complex was found to be larger in solvents with higher Gutmann donor number. The exchange rates and thermodynamic parameters of lithium cryptate exchange in various solvents were determined from the 7Li NMR tempera- ture-dependent data. Hourdakis and Popov (80) have used 7Li, 23Na, and 133Cs NMR to study alkali complexes with cryptand C222—dilactam in various solvents. They found out that the complexing ability of the dilactam is similar to, but weaker than, that of the cryptand C222. Recently, Smetana and Popov (81) have studied complexes of Li+ ion with several crowns in various solvents using 7Li NMR. Sodium 23 NMR measurements have been obtained on many antibiotic ionophores in chloroform and methanol solutions (82). In all cases, addition of ionophores to the sodium 37 ion solution broadens the 23Na resonance lines. Despite the similar nature of the complexes, the 23Na chemical shifts were found to be very different for different anti- biotics. The complexation of Na+ ion with pentamethylene tetrazole in nitromethane has also been studied by 23Na NMR (83). Addition of the crowns, such as 18-crown—6 derivatives, to the sodium salt solutions of various solvents (84,85) has been shown to result in a very appreciable broadening Of the sodium-23 resonance so that the resonance line could not be detected. This is because most crown ethers tend to form two-dimensional complexes with the alkali ions which could distort the spherically symmetrical electric field around the solvated sodium ion and, there- fore, broaden the 23Na resonance line. GrandJean 22.21: (86) have determined the thermo- dynamic parameters for complexation of a heptadentate non- cyclic crown ether with Na+ ion in pyridine by 23Na-NMR spectroscopy. Using 23Na resonance line broadening measure- ments, they have determined the relaxation rate, and by measurement of relaxation rate as a function of composition have calculated complex formation constants at different temperatures. From the resulted thermodynamic parameters they have concluded that the complex has a "wrap around" structure. Sodium-23 NMR has been extensively used to study the 38 exchange kinetics of Na+ ion with crowns (87,88) and cryp- tand (89,91) in different solvents. Shchori g£_al, (87,88) have investigated the kinetics of Na+ complexes of dicyclohexyl-lB-crown-6 and dibenzo-l8-crown-6 and its derivatives in various solvents. The life times of free and complexed sodium ion and the pseudo first-order rate constant for the decomplexation reaction have been found from line shape analysis as a function of temperature. Different substituent groups on the ligands had a sig- nificant effect on the decomplexation reaction. Dye and co-workers (89,91) Obtained two 23Na resonance signals for Na+-C222 cryptate solutions with the excess of the sodium salt in various solvents. One resonance corresponds to the sodium ion inside the cryptand cavity, and the other corresponds to the uncomplexed solvated sodium ion. The rate constants, activation energies and thermodynamic parameters for the decomplexation reaction were obtained from line shape analysis Of the 23Na NMR temperature-dependent data. Shih and Popov (92) studied complexation reaction between K+ ion and several crowns and cryptands in various non-aqueous solvents by 39K NMR spectroscopy. They found evidence for formation of an inclusive complex between K+ ion cryptand C222 but an exclusive one for K+-C221 cryptate in solution. The K+-18-crown-6 complexes were found to be quite stable in non-aqueous solvents. The lS-crown-S's 39 were found to form both 1:1 and 2:1 sandwich complexes with K+ in all non-aqueous solvents used. Shporer and Luz (93) studied the longitudinal relaxa- tion time T1 of the potassium-39 nucleus as a function of temperature in methanol solutions in the presence of di- benzo-18-crown-6 using 39K-NMR technique. The rate of the decomplexation reaction and the activation energy for the reaction were then calculated from the resulting data. Popov and co-workers (63,96,94-96) have obtained interesting results from a cesium-133 NMR study of both the kinetics and thermodynamics of crowns and cryptands complexes with Cs+ ion in non-aqueous solvents. From 13303 chemical shift measurement as a function of ligand to metal ion mole ratio, they have obtained evidence of a two-step complexation reaction between Cs+ ion and 18— crown-6's ligands (63,64,94). The formation of a 1:1 complex is followed by the addition of a second molecule of crown to give a 2:1 sandwich complex. The thermodynamic and kinetic parameters for the two step reaction of Cs+ ion with 18-crown—6 in pyridine have been obtained from the mole ratio data at different temperatures (64). The chemical shift of the 1330s resonance was studied as a function of cryptand C222 to Cs+ at various tempera- tures in different solvents (95). The formation constants for Cs+-C222 cryptate were then determined from the resulting data. The limiting chemical shifts for Cs+-0222 40 complex in dimethylformamide, propylenecarbonate and acetone were found to approach to the same value at about -100°C. From these results the authors suggested that the complex is exclusive at higher temperatures but inclusive at lower temperatures. The thermodynamic parameters were calculated by analysis of the data Obtained at different temperatures. Thallium-205 NMR has been used to study the metal ion binding to biological macromolecules such as pyruvate kinase (97). It also has been used to investigate com- plexation of T1+ ion with crowns (98) and antibiotics (99). 1.3. CONCLUSIONS From the above discussion, it is evident that multi- nuclear NMR provides a very powerful tool for studies of complexation reactions in solution. The full potential of this method has not been realized; however, with con- tinuous improvement in NMR instrumentation its increasing usefulness for different kinds of studies can be predicted. Information on the thermodynamics and kinetics of complexa- tion reaction can also be obtained by this method. The subject of this thesis is a multinuclear NMR study of thermo- dynamics and kinetics of large crowns in various non- aqueous solutions. CHAPTER 2 EXPERIMENTAL PART 41 2.1. SYNTHESIS AND PURIFICATIONS OF LIGANDS 2.1.1. Synthesis of Dibenzo-30-Crown-10 The method for synthesis of dibenzo-30-crown-10 was essentially based on Pedersen's (1) published method for the synthesis of benzo-lS-crown-S. By reacting catechol with l,ll—dichloro-3,6,9-trioxaundecane both benzo-15- crown—5 and dibenzo-30-crown-10 are formed. The main product of the reaction is benzo-15-crown-5. The procedure is slightly modified for separation of dibenzo-30-crown-10 from benzo-15—crown-5. Catechol (88 g, 0.8 mole) was dissolved in 800 ml Of dried n-butanol containing 0.8 mole of KOH (53 g, 85% pure) in a 2.5 liter three necked flask, equipped with a thermometer, a dropping funnel, and a condenser. The equipment was under a nitrogen atmosphere. The mixture was heated to reflux, when solution became muddy-white- greenish in color. l,1l-Dich1oro-3,6,9-trioxaundecane (93 g, 0.4 mole), prepared according to Pedersen's method (1), was then added dropwise over four hours with continued heating, and the mixture was further heated for an additional three hours. The mixture was then cooled to room temperature, and 0.8 mole of KOH (53 g, 85% pure) was added. The mixture was again heated to reflux. Another 0.4 mole (93 g) of 1,11- dichloro-3,6,9-trioxaundecane was then added dropwise 42 43 over four hours. The mixture was refluxed for about 24 hours, while the color of the solution changes to dark brown, and then was cooled and evaporated under vacuum in a rotO-vac apparatus. After cooling the mixture, a solution of 20 ml of concentrated HCl in 150 ml water and 500 ml of chloroform was added. The chloroform layer (lower layer) was then separated and was washed three times with 400 ml of each Of saturated NaCl-50% NaOH solutions, and once with 400 m1 of saturated NaCl solution (separation of phases was slow). The chloroform solution was then dried Over Nazsou. The chloroform was then evaporated using a rote-vac. A soft solid was then obtained which was suspended in a warm (80—90°C) n-heptane solution in a continuous liquid- liquid extractor for several days. After cooling the yellow n-heptane extract to about 40-50°C, two layers were obtained. The upper layer containing benzo-lS-crown- 5 was separated. The lower layer was recrystalized from pure acetone to give white plate crystals of dibenzo-30- crown-10 with a melting point of 106.5°C. The yield of reaction for benzo-lS-crown-S and dibenzo-30-crown-10 were found to be 37% and 3%, respectively. 2.1.2. Purifications of Ligands Some dibenzo-BO-crown-lo was Obtained through the courtesy of DuPont company. The ligand was recrystallized 44 from acetone and vacuum dried. Dibenzo-2l—crown-7 and dibenzo-24-crown-8 (Parish Chemical Company) were re- crystallized from n-heptane and dried under vacuum for three days. The melting points of well defined crystals were found to be 107 and 103°C, respectively, which are the same as the reported values (17). 1,10-Diaza-18- crown-6 (Merck Company) was recrystallized from reagent grade n-heptane and dried under vacuum for several days. Benzo-lS-crown-S was recrystallized from pure n-heptane and vacuum dried for 72 hours. 2.2. SOLVENTS AND SALTS 2.2.1. Solvents Reagent grade acetone (Mallinkrodt) was refluxed over calcium sulfate, fractionally distilled, and dried over 4A Linde molecular sieves. Spectrophotometric grade nitro- methane (Aldrich) was refluxed over night over phosphorus pentoxide, fractionally distilled, and dried over activated molecular sieves for 24 hours. Dimethylsulfoxide (Fisher) was dried over Linde 4A molecular sieves and vacuum dis- tilled. Dimethylformamide (Fisher) was vacuum distilled over phosphorus pentoxide. Propylene carbonate (Aldrich) was dried for 48 hours over Linde 4A molecular sieves followed by vacuum distillation. Acetonitrile (Mallin- krodt) was refluxed over calcium hydride, fractionally 45 distilled and dried over molecular sieves. Trimethylene oxide (Aldrich) was dried over activated molecular sieves for 48 hours. Tetramethylguanidine (Eastman) was refluxed over granulated barium oxide and fractionally distilled. The molecular sieves used were activated by heating them at 600°C for 12 hours under a nitrogen atmosphere. Analyses for water in solvents were carried out with an automatic Karl Fischer Aquatest (Photovolt Corp.) titrator. The water content Of the sOlvents after drying was found to be less than 100 ppm. 2.2.2. Salts Lithium perchlorate (Fisher) was dried at 190°C for several days. Sodium tetraphenylborate (J. T. Baker) was dried under vacuum at 60°C for 72 hours. Potassium hexafluorophosphate (Pfaltz & Bauer) was purified by recrystallization from water and dried under the vacuum at 110°C for 72 hours. Cesium thiocyanate (Pfaltz & Bauer) was recrystallized from reagent grade methanol and dried under vacuum for several days. Cesium tetraphenylborate was prepared by mixing tetrahydrofuran solution of sodium tetraphenylborate with equimolar amount Of concentrated aqueous solution of cesium chloride (Ventron Alfa Product). The resulting fine white precipitate was collected, washed several times to remove any adhering sodium salt and dried under vacuum at 70°C for 72 hours. Thallium (I) 46 perchlorate (K & K) was recrystallized from water and dried at 110°C for 24 hours. 2.3. SAMPLE PREPARATION All solutions were prepared in a dry-box under a nitro- gen atmosphere to maintain their water content at the lowest possible level. Dilute solutions of the salts were prepared by appropriate dilution of a stock solution. The ligand solutions were prepared by proper dilution of a stock solution (if the ligand was soluble enough) or by weighing out in the desired amount into a 2 ml volumetric flask followed by dilution with the solvent or the solution. 2.4. INSTRUMENTAL MEASUREMENTS Lithium-7, sodium-23, cesium-133, and thallium-205 NMR measurements were carried out on a modified Varian Associate DA-60 spectrometer equipped with a wide-band probe capable of multinuclear operation (101), Operating at a field of 14.09 kgauss in a pulsed Fourier transform mode. The spectrometer is equipped with an external proton lock to maintain the field stability. The NMR spectrometer is interfaced to a Nicolet 1083 computer for pulse genera- tion, data collection, and data treatment. A previously described program (102) was used for pulse generation and collection of the resultant free induction decay (FID) 47 signal. Data treatment was-performed using the Nicolet FT-NMRD program (NIC-80/S-7202-D) (103). The experimental conditions are given in Table I. Table I. Experimental Conditions for Metal Ion NMR. Resonance External Reference Nucleus Frequency (MHz) Solution 7L1 23.32 4.0 M LiClOLl in H20 23Na 15.87 3.0 M NaCl in H20 133Cs 7.87 0.5 M CsBr in H20 205Tl 34.61 0.3 M T1NO3 in 320 All the chemical shifts for sodium-23, cesium-133, and thallium-205 reported in this thesis are referred to infinitely dilute aqueous Na+, Cs+, and Tl+ solutions, and the chemical shifts for lithium-7 are referred to a 4.0 M LiClOu aqueous solution. Downfield (paramagnetic) chemical shifts from the reference are indicated as ne a- tlzg, In order to keep the chemical shift of external reference constant, an insulated reference tube (64) was used in the measurements of cesium-133 Chemical shifts as a function of temperature. The reported chemical shifts are also corrected for the differences in the bulk diamagnetic susceptibility 48 between the sample (nonaqueous) and the reference (aqueous) solutions according to the Live and Chan (104) equation for non-superconducting spectrometers sample) V _ 2n ref 6corr " 6obs + I? (XV - X where xgef and Xsample are the volume susceptibility of the reference and the sample solutions, respectively. The magnitude of correCtions, calculated on the basis of the published susceptibilities (105), and the physical properties (106,107) for the solvents used in this study are shown in Table 6. Carbon-13 NMR spectra were obtained on a Varian CRT-20 spectrometer operating at a field of 18.68 kgauss in a pulsed Fourier transform mode. Acetone was used as an external reference and D20 was used to lock the system. All carbon-l3 chemical shifts are reported with respect to TMS. 2.5. DATA HANDLING The complex formation constants were obtained by fitting the chemical shift-mole ratio data to appropriate equations (which will be discussed in detail later) using the least squares program KINFIT (108) on a CDC-6500 computer. A linear least squares program was used to obtain enthalpies and entropies. 49 .AAOHV oosooooomo . Amoav mocmhmhmmn .Aooao oocohooomo NAN.oI omm.o IIII o.HH ocaoacosmflsoooERAooB omm.oI «Ho.o H.mm a.mH osfioaa:a ooH.oI amo.o H.mH o.mo ooocooooo ocoa:aoto amo.oI Hmm.o s.m m.mm ococooancHz mm:.oI omm.o os.mm s.mm Hocoooo: H:N.OI omo.o m.mm s.oe ooaxooasnflsoooeHo mom.oI oom.o o.om s.om ooaeoewooaacooeHo omm.oI mmm.o H.3H m.om oafiooasooooa mam.oI ooa.o o.sH s.om osoooo< “EaavomIHom so coapooppoo n:flflwmmhwmmuwsm hocoa ofinuooamfim .omIaom mom .m manna CHAPTER 3 MULTINUCLEAR NMR STUDY OF DIBENZO—30-CROWN-10, DIBENZO-24-CROWN-8, AND DIBENZO-21-CROWN—7 + COMPLEXES WITH Na+, K , 08+, and T1+ IONS IN NONAQUEOUS SOLVENTS 50 3.1. INTRODUCTION Previous studies in our laboratories (109—112) and elsewhere (113—117) have shown that the nuclear magnetic resonance Of thallium and alkali nuclei Offers a very sensitive technique for the studies of changes in the im- mediate chemical environment of the thallium and alkali ions in solution. The chemical shifts and line width of the resonances can given information about ion-ion, ion— solvent, and ion-ligand interactions. During the past decade alkali metal NMR has been used extensively to study the thermodynamics and kinetics of the complexation reac- tion between alkali metal ions and crowns and cryptands (109-118). Among the crown ethers, large molecules such as dibenzo- 30-crown-10 and dibenzo-24-crown—8 have some interesting properties. These are very flexible molecules with enough oxygen atoms in the ring so that they can twist around a metal ion with a suitable size to envelope it completely and form a three-dimensional "wrap around" complex (26). Alkali complexes of dibenzo-30-crown-10, dibenzo-24- crown-8, and dibenzo-21-crown-7 have been studied by potentiometry (4,31), polarography (35), calorimetry (34), spectrometry (32), and proton and carbon-13 NMR (33) in different solvents. The purpose of the study described +, 03+, and Tl+ in this chapter was to investigate Na+, K ion complexes of the above mentioned ligands in a number of nonaqueous solutions by the multinuclear NMR technique. 51 52 3.2. COMPLEXATION OF Na+, K+, AND CS+ IONS WITH DIBENZO- 30-CROWN-10 3.2.1. 0330010 Complexes with Na+ and K+ Sodium-23 chemical shifts were determined as a function of dibenzo-30-crown-lO/sodium ion mole ratios in nitro- methane, acetonitrile, and pyridine solutions. The result- ing data are given in Table 7 and the mole ratio plots are shown in Figure 3. In all cases only one population- average resonance was observed indicating that the ex— change of the metal ion between the bulk solution and the complex is faster than the NMR time scales. In the case of pyridine solutions, the addition of the ligand to the Na+ solution produces a gradual diamagnetic shift Of sodium- 23 resonance which begins to level off at a mole ratio of about 1, which indicates the formation Of a 1:1 complex of the sodium ion with the ligand. On the other hand, in the case of nitromethane and acetonitrile solutions, the chemical shift 1g. mole ratio plots show three distinct inflection points at the ligand/ metal mole ratios of about 0.5, 0.7, and 1 indicating the formation of three complexes with the respective stoi- chiometries DB30010-2Na+, 2DB30C10-3Na+, and DB3OClO-Na+. The synthesis and isolation of crystalline 1:2 and 1:1 (ligand to metal ion) complexes of sodium tetraphenyl— borate and dibenzo—30-crown-10 have been previously 53 msa 0m.0H m=.m 00H 0H.0H m0.m mm: no.0: mm.H mm: mm.0a 0a.: m0 m0.0 mm.H 00H m0.0H 00.0 mm No.0 0m.a mm: 00.0: 00.0 0mm ma.» 0H.m mm 00.0 0a.: 00m mo.HH 0s.0 mom mm.s e0.H 00 se.0 00.0 H00 HA.HH 0s.0 0mm 00.0 00.: mo oa.0 00.0 mam H0.HH 00.0 m0m ms.0 0H.H 0s 00.A H0.0 we: 00.HH 00.0 00H m0.0 00.0 mm ss.s as.0 00m 0m.0 sm.0 Hod mH.m mw.o a» mo.m Hm.o 23H H~.m mm.o 00H Hm.m 00.0 as 0H.0 00.0 mm: ss.0H mm.0 AHA ss.a 0m.0 m0 00.s 0m.0 0H: m0.HH mm.0 mm ms.0 0H.0 mm sm.s 0H.0 00 Hm.NH 0H.0 a: ss.0I 00.0 HH m0.e 00.0 0: 00.0: 00.0 AN:VN\H>0 Asaave +:\o AN:VN\H>< Asaave +o2\o Asmvmxaao Asaave +o2\o . as 2: :z I .ooom no muco>aom macano> CH ouononamcoca Iosooa season :00.0 can: aoonano 0HIcsos0I0mIoscooao 0o soaom capo: oHo: .s oases 54 IS: A Io - . ‘r i 8 8 (ppm) M I C 5 .. O _ 1 1 l I I l (15 L0 L5 2£> 25 [DB SOCIO]/[No+] Figure 3. Sodium-23 Chemical Shifts Kg. [DB3OClO]/[Na+] Mole Ratio in Different Solvents. A-Nitromethane, B-Acetonitrile, C-Pyridine. 55 reported by Truter and coworkers (25). These three species were isolated from acetonitrile solutions in the form of well defined crystals. The melting points of 1:1 and 1:2 (ligand to metal ion) complexes were found to be very close to the reported values. The crystalline structure of the third species which we supposed to be 2:3 complex were determined. The results showed that it is not the 2:3 species, but instead it is the 1:2 complex associated with four molecules of acetbnitrile of crystallization. It seems that 2:3 complex is quite unstable and exists only in solution. It has been shown previously that carbon—l3 chemical shifts Of the carbons in the ether region of cyclic poly- ethers are sensitive to the conformational change of the ligand upon complexation (119). In order to get further information about dibenzo-30-crown-10 interactions with the sodium ion, we studied the chemical shift of the polyether chain carbons as a function of sodium and potassium ion concentrations relative to the concentration of the ligand. The results, obtained in nitromethane solutions, are shown in Figures 4 and 5. In the case of the DB3OCIO-K+ ion system. the addition of potassium hexafluorophosphate to a DB3OC10 solution results in a gradual coalescence of the four resonances and only one signal is obtained at equimolar concentrations of the potassium ion and the ligand. This behavior seems 56 R: 25,, All. 11 , l 21) -quJ LO 1 [J I 73 7| 69 67 73716967 8 Figure 4. Carbon-13 Chemical Shifts at Various [Metal IonJ/ [DB3OC10] Mole Ratios (MR) in Nitromethane. A-Sodium Ion at MR=0.0, 1.0, and 2.0; B—Potassium Ion at MR-0.0, 0.5, and 1.0. 56 5:5 , 1 2;) L] Aaequ I!) J l l l l l 73 7| 69 67 73 7l6967 A B Figure u. Carbon-13 Chemical Shifts at Various [Metal IonJ/ [DB30C10] Mole Ratios (MB) in Nitromethane. A-Sodium Ion at MR=0.0, 1.0, and 2.0; B—Potassium Ion at MR-0.0, 0.5, and 1.0. 57 72- ’OC3 C4 < ~—IC2 69- 68'- CI l l l l I l (15 L0 L5 2£> 25 30 [No+]/ [083000] Figure 5. Carbon-13 Chemical Shifts for the Four Poly- ether Chain Carbon Atoms at Various [Na+]/ [DB3OClO] Mole Ratios in Nitromethane. 58 to indicate an essentially equal interaction between the ten oxygens of the polyether ring and the potassium ion. Naturally, such equal interactions is only possible if in solutions the ligand is "wrapped around" the cation as postulated by Live and Chan (33) in solution and shown by Truter and coworkers (26) for the solid complex. Considerably different behavior is observed in the case of the sodium ion. The details are given in Figure 5. Between mole ratios of 0.0 and 2.0 all four carbons behave quite differently. While the initial addition of the sodium ion results in a chemical shift of carbons 2, 3, and H, the chemical shift of carbon 1 remains unaffected. 0n the other hand, between mole ratio of 1.0 and 2.0 the chemical shifts of carbons 2 and 3 remains constant while carbon N and especially carbon 1, show a significant down- field shift. Beyond mole ratio of 2.0, the resonance fre- quencies are constant. The results show that the addition of the sodium ion to DB3OC10 results in at least two conformational changes of the ligand molecule following the formation of DB3OC10°Na+ and DB3OC10'2Na+ complexes. No evidence for the forma- tion of the 2DB30010-3Na+ complex was observed; however, it is to be expected that the 23Na chemical shift is a much more sensitive probe of the sodium ion interaction than the 13C chemical shift. Once again, these data support the conclusions of Live and Chan (33) that the 59 DB30ClO'Na+ complex has a different configuration from the DB3OC10 complex with potassium ion (and presumably with 2+ ions). the Cs+ and Ba An attempt was made to calculate the formation con- stants of DB3OC10 complexes with the potassium ion from the variation of the carbon-13 chemical shift as a func- tion of K+/ligand mole ratio. It was found that in nitro- methane and acetonitrile solutions log Kf was greater than 5, while in acetone solutions log Kf = “.3 i 0.1. 3.2.2. 0830010 Complexes with Cs+ The variation of cesium-133 chemical shift as a func- tion of the ligand/Cs+ mole ratio in nitromethane, aceto- nitrile, acetone, methanol, and pyridine solutions at 30°C are shown in Figure 6 (the data are given in Table 8). It is seen that the shift is diamagnetic in acetonitrile solutions and paramagnetic in all others. The shifts begin to level off at a mole ratio of about 1, indicating the formation of a 1:1 complex. It is interesting to note that the limiting chemical shifts for the complexed cesium tend to approach each other indicating that in the complex the cation is largely insulated from the solution and, once more, confirming the "wrap around" structure. The formation constants for the DB3OClO°Cs+ complex in different solvents were determined from the variations of the 133Cs chemical shifts with the ligand/Cs+ mole 60 .encmmoc .ZQmmoc .eoflomoe me.HH mH.m mm.m oo.m mm.mH mo.m m:.HH es.fi 3H.ea ee.a mm.mm mm.H :m.w om.H oo.mH ow.H mm.HH me.a :H.ea om.H ma.om mm.H mm.m mm.H HH.mH me.a mm.HH NN.H Hm.ea em.H mm.mm mm.H NH.: mH.H oe.mH m=.H NH.NH 0H.H mm.ma mm.H Hm.mm mm.H me.m NH.H o=.mH om.H mm.me mo.H mH.mH eo.H Hm.em eo.H mm.oa mm.o H=.om mH.H Hm.=a mm.o ee.om mm.o oo.mm mm.o mm.mu Hw.o oo.mm mm.o mm.mH om.o em.mm mm.o oe.mm mm.o ee.mn ce.o oe.mm em.o m=.ma mm.o :~.mm om.o Hu.om m>.o um.mH| mm.o Hm.om an.o mm.=m e=.o No.0m mm.o HH.mm em.o om.mau om.o ee.mm om.o m:.mm mm.o m~.mm mm.o Hm.mm om.o Hm.mmn mm.o :.mm mm.o Hm.mm oo.o oo.me oo.o Ho.oe 00.0 mm.mma oo.o He.Hm oo.o Aeccve +mo\q Asccve +mo\a Aaccve +mo\q Aeccve moxq Ashave +mo\q hmocz ce< l .0 com um mpCo>aom msofihm> ce eOH mo Emoo.o nee: mcxcaceOo afiuceoeouomucuechen cc secem cheem mac: .m cane + 61 50 3O 8 (ppm) I 0% To _ IJS— 2.0 2.5 [DBBOCIO]/[Cs*] Figure 6. Cesium-133 Chemical Shifts E. [DBBOClO]/[Cs+] Mole Ratio in Different Solvents. A-Nitromethane, B—Methanol, C-Acetone, D-Pyridine, E-Acetonitrile. 62 ratio according to the following method. Assuming that only cation-ligand interactions are important, for the two chemical environments of the metal ion, freely solvated and complexed, in which the exchange rate is fast as com- pared to the NMR time scale, a population average shift is observed 6obs = XMGM + XMLGML (1) where Gob is the observed chemical shift in ppm, XM S and XML are the respective chemical shifts for the two sites. Equation 2 is easily derived from Equation 1 where O 6 =fl obs t (5M ‘ 6ML) + 6ML (2) O 3 CM is the concentration of free metal ion and 0% is the total concentration of metal ion. Assuming formation of a 1:1 complex, we have the equilibrium where L is the ligand. The concentration equilibrium constant can be written as 63 where CML is the concentration of the complex, CM is the concentration of free metal ion, and CL is the concentra- tion of free ligand. By a simple algebraic substitution, Equation 3 can be written as t CM ‘ CM Kf g t t (u) where CE is the total concentration of the ligand. From Equations 2, 3, and u we can obtain Equation 5 which relates the observed chemical shift to the formation constant, the total concentrations of the metal ion and the ligand (0; and 0E respectively), the limiting chemical shift of the complexes metal ion (GML), and to the chemical shift of the free metal ion (6M). 2 2 _ t t 2 t 2 t 2 t t Gob, - [(KfcM - Kch - 1),: (KfCL + KfCM - 2KfC CL + 2K 0t + 2K 0t + 1)1/2].(Efl:ifl£. + 6 ( ) r L r M t ) ML 5 2KfCM In Equation 5, 0; and CE are known, 6M can be easily determined from the measurement on solutions of free metal ion. The equation then contains two unknowns Kr and 5ML' t t The procedure then is to input the Sobs’ CM’ CL’ and 5M parameters and vary Kr and 6ML using a computer least 6“ squares program KINFIT (108). The procedure will continue until the calculated chemical shifts correspond with the experimental chemical shifts within the error limits. A typical fit of the data for DB30C10-Cs+ in methanol at 30°C is shown in Figure 7. The results for the DB3OC10°Cs+ complexes in various solvents are shown in Table 9. It is seen that in acetone solutions the stability of the complex is not affected by a change in the concentration of salt or of the counter ion. It is evident, therefore, that at low concentrations of the cesium salts, which we used, the formation of the complex is unaffected by ion pairing. It is reasonable to assume that the same situation will exist in solvents with higher donicities and/or higher dielectric constants such as nitromethane, acetonitrile, and methanol. Comparison of our values with those reported in the literature (and obtained by different techniques) show a satisfactory agreement (Table 9). Since the cesium ion is rather weakly solvated be- cause of low charge density of the cation, it is not surprising that the stability of the complex is only marginally dependent on the nature of the solvents. 65 .00 em up Hosanna: ca +eonoaoomma ace memo caeem mac: mmaueeamco can no can handmade .e madman padz. do oz on tour o-uon mu mam—m:<1m ocoooooooooooooooouooooo oz Hm----m---'m----m'-'-m---'m---'m----m----m-'--w----m-"-m----m'---m'---m-'--m----w----m----w----r-'--_ W a I u . ~ n u w n r _ i u o 1 i. . - _. u . ._. .. a u u w _. a K m c . 11 fi . _. a T .T w . m. a w e i X a m .— * i n _ .. w ._. — w 2 —w.----m'--'m---'m----m---'m----m----m----mal'|'ml---w-lu-'m----m---I'm----m---'m----m----m"--m---'m---. > (pawn >6 x (pawn U242o. pz_od oup<42uqcu a mzoN°oU9€¢o n QCh Ulb h< UDJ¢> canmma Q0 Q—oNox Wu 4(U~b¢w> Qciuhcho u hZUZUCUZ—oNOIwOOKo u hIO—m UIb h< u34<>o cc "bum; Urn h¢ NDJ<>oA—0¢(>¥~X~RO .WWWPM¢mwmmmmwwm 66 Table 9. Formation Constants and the Limiting Chemical Shifts of DB3OClO-Cs+ Complexes in Various Solvents. Solvent log Kf 611m(ppm) Nitromethane H.30i0.05 18.5810.0h Acetonitrile 3.39:0.09 13.5630.95 Acetonitrile 3.50a ---- Acetone - 3.99:0.08b lu.69:o.01 Acetone u.ouio.05° 15.05i0.02 Acetone 3.96:0.07d 15.89ao.o7 Acetone 4.05:0.06e 15.28i0.05 Acetone 14.23f ---- Methanol “.18i0.07 l6.SOiO.ll Methanol 14.23g ---- Pyridine “.4120.10 11.3510.02 aReference 35. bo.oosl~_a_0ssCN. °o.oosg Cs Picrate. d0.005fl’CsBPhu. e0.0025M|CsBPhu. fReference 33. gReference A. 67 3.3. COMPLEXATION 0F Na+, Cs+, and Tl+ IONS WITH DIBENZO- 2U-CROWN-8 3.3.1. 032u08 Complexes with Na+ The complexation between sodium ion and dibenzo-2u- crown-8 was studied in nitromethane, acetonitrile, dimethyl— formamide, dimethylsulfoxide, and pyridine. The measured sodium-23 chemical shifts at various ligand to sodium ion mole ratios and the line widths of the resonance lines at the half height at 30°C are listed in Table 10. The mole ratio plots are shown in Figure 8. It is seen im- mediately that the solvent plays an important role in the complexation reaction. In solvents of high donicity, such as DMF and DMSO, the sodium-23 resonance is almost independent of the ligand/sodium ion mole ratio. This behavior indicates that in these solvents the immediate environment of the sodium ion is not changed upon the addition of the ligand indicating formation of a very weak complex at the best. On the other hand, in nitromethane, acetonitrile, and pyridine solutions the sodium-23 resonances shift upfield or downfield with some indication of a break at a mole ratio of about 1 implying the formation of a 1:1 DB21408-Cs+ complex. The formation constants and the limiting chemical shifts of the complex in different solvents are given in Table 11. It is obvious that in solvents with low solvating \ 68 a: om.m m:.m um mm.» wm.m w: mm.= mm.m mm mm.m mm.a mm pm.» uo.m u: mm.= ze.H mm um.m mm.H um um.» mm.H m: om.: mm.a m: mm.m mm.H mm mm.~ :N.H m: m~.= NH.H o: ua.w oa.a am om.m mo.H m: m~.: :o.a mm Ha.m mm.o mm mm.m mm.o o: m>.: No.0 mm wo.m mw.o mm mm.m mm.o H: mm.: m~.o am oo.m m~.o om ma.m m~.o 0: mm.: mm.o ma mm.» mm.o mm m>.aH mm.o mm on.: 00.0 Ha om.» oo.o NH no.=a oo.o “Naom\aaa Aeaove +e2\a Asavm\aea Assays +e2\o Assvm\aee “eases +c2\a mic z< 22 I. .ooom pm mpcm>Hom msoaum> CH mpmnonazcona Imnpoe Esaoom Emmo.o 20H: moonQEoo mlczopolamlouconwa mo madam oaumm oaoz .oa manna 69 MHH mm.m Hm.m mm mm.o ma.m NHH ms.m mm.a mm e=.o mo.m OHH mm.m mm.H mm «3.0 Ho.H mos om.s om.H a: am.o NN.H mos oe.e OH.H om am.o oa.a sea om.» sm.o Hm He.o om.o em mm.m mm.o Hm mm.o mm.o we no.0 ma.o om wm.o ma.o om mH.m em.o o: om.o om.o om mH.H- oo.o we HH.o oo.o Anmvm\aoa Aeoove +ezxq Anmvm\aa< “sodas +ozxq . ad omzo .ocssaesoo .OH canoe 7O 14 12 - 10 - PY AN 8" ‘nur sot i;._ a_4n 5 NM (ppm 6- DMF _.__ 4.. 2.. DMSO __.___.__._'_______.__ ‘fit— 0“ °7 J l l I 11 0.0 0.5 1.0 1.5 2.0 2.5 [DB24C8]/[No+] Figure 8. Sodium—23 Chemical Shifts y_s__. [DB24C8j/[Na+] Mole Ratio in Different Solvents. 71 Table 11. Formation Constants and the Limiting Chemical Shifts of DB2hCB-Na+ Complexes in Various Solvents. Solvent log Kf 611m(ppm) Nitromethane 3.74:0.12 7.82:0.01 Acetonitrile 2.95:0.07 8.33:0.01 DMF m0 -_-- DMSO - m0 ---- Pyridine 2.89:0.10 9.20:0.09 72 abilities the complex is more stable. Once again pyridine seems to be an exceptional case. The limiting chemical shifts of the complexed sodium ion in nitromethane, acetonitrile, and pyridine solutions are close to each other indicating that in these solvents the complexed sodium ion is no longer exposed to the sol- vent molecules. This is only possible if the sodium ion is located inside the cavity created by the twisting of the ligand around the Cation. Another evidence which sup- ports the formation of such a "wrap around" complex between DB2HC8 and the sodium ion is the existence of the rela- tively narrow signals for sodium-23 resonance in all solu- tions. The width of the 23Na resonance lines in this case are less than one half of those observed for Na+-DB30C10 system in the same solvents (Table 7). For nuclei such as 23Na with appreciable quadrupole moment, we expect to have a broader resonance line for a more unsymmetrical environment around the nucleus. Thus existence of rela- tively narrow 23Na resonance lines in this case indicates a symmetrical environment around the sodium ion which can only be obtained by the formation of a three-dimensional complex. 3.3.2. DB2HC8 Complexes with Cs+ Cesium-133 chemical shifts were determined as a func- tion of the ligand to the cesium ion mole ratio. The data 73 are given in Table 12 and the mole ratio plots are shown in Figure 9. In all cases, the cesium-133 resonance has a gradual shift, paramagnetic or diamagnetic depending on the nature of the solvent, until a mole ratio of about 1 is reached and then begins to level off. This behavior indicates formation of a 1:1 DB2u08-Cs+ complex in all solvents. The formation constants and the limiting chemical shifts of the complex in various solvents are shown in Table 13. From the results it is obvious that solvent plays an important role in the complexation. In strong solvating solvents such as DMF and DMSO the complex is much weaker than in solvents of low and medium donor strength such as nitromethane, acetonitrile and acetone. This behavior shows the existence of a competition between the ligand and the solvent molecules for the cesium ion. The only ex- ception is the case of pyridine where despite the high Gutmann donor number of the solvent the complex is un- expectedly stable. This exception probably results from the relatively weak interaction between the "soft base" nitrogen atom of pyridine and a "hard acid" (cesium ion). Three different cesium salts at different concentra- tions were used to investigate the effects of the anion and of the concentration on the complexation of cesium ion with dibenzo-24-crown-8 in acetonitrile solution. The results are also given in Table 13. As seen in all 7h Table 12. Mole Ratio Study of Dibenzo-2u—Crown-8 Complexes with 0.005flCs+ Ion in Various Solvents at 30°C. NMa ANa ACa MeCHa + + + + L/Cs 6(ppm) L/Cs 6(ppm) L/Cs 6(ppm) L/Cs 6(ppm) 0.00 57.57 0.00 -33.55 0.00 19.27 0.00 u6.15 0.21 53.31 0.25 -23.2u 0.25 21.19 0.20 nu.76 0.60 u5.62 0.62 -7.50 0.60 23.82 0.5a A1.50 0.70 uu.38 0.72 -3.70 0.77 25.15 0.78 u0.03 0.9a 91.37 0.8u -0.20 0.98 26.69 0.9a 39.02 1.05 no.59 0.98 5.85 1.12 27.70 1.02 38.63 1.10 u0.u2 1.09 8.32 1.25 28.16 1.18 37.87 1.18 39.86 1.13 8.55 1.81 28.u8 1.3u 37.39 l.u0 39.0u 1.38 12.12 2.19 28.79 1.60 37.09 1.87 38.56 1.78 13.68 2.66 29.02 1.76 36.92 2.39 38.u2 2.32 lu.1u 29.0u 2.07 36.79 2.88 38.u2 2.78 1M 22 2.78 36.70 DMFa DMSOa Pyb + + + L/Cs 6(ppm) L/Cs 6(ppm) L/Cs 6(ppm) 0.00 0.26 0.00 -67.26 0.00 23.99 0.3a 3.12 0.u2 —63.27 0.38 22.82 0.60 5.23 0.71 -60.63 0.56 22.uu 0.80 6.68 0.86 -59.02 0.83 21.75 0.8a 6.86 1.00 -58.53 0.92 21.36 0.96 7.78 1.1u -57.15 1.05 21.19 1.12 8.78 1.20 -57.29 1.26 20.98 1.30 9.71 1.38 -55.90 1.52 20.88 1.10 10.11 1.85 -52.79 2.03 20.81 1.85 12.19 2.30 -u9.92 2.39 20.79 2.60 19.20 2.80 -u7.52 2.83 20.79 aCsSCN. b CSBPhu . 75 50b I l I l J J 0.0 0.5 3.0 2.5 3.0 1.0 1.5 [DBZAC8]/[C 5* + Figure 9. Cesium—133 Chemical Shifts 1g. [DB214C8J/[Cs J Mole Ratio in Different Solvents. 76 Table 13. Formation Constants and the Limiting Chemical Shifts of DB214CBoCs+ Complexes in Various Solvents. Solvent Log Kf 611m(ppm) Nitromethane M.lli0.08 38.2610.03 Acetonitrile 3.9uao.07a lu.86:0.02 Acetonitrile 3.89:0.07b lu.5010.07 Acetonitrile 3.90:0.011c lh.8210.06 Acetonitrile 3.97:0.011d 19.2310.06 Acetone 3.71:0.09 29.2“10.06 Methanol 3.65:0.05 36.h710.03 Methanol 3.78:0.08e ---- Dimethylformamide 2.10:0.0“ 2U.9330.96 Dimethylsulfoxide 1.61:0.0M -8.32i3.h6 Pyridine 9.00:0.03 20.7510.07 a0.00SM_CsSCN. b0.0101310sSCN. c0.005M CsBPhu. d0.005110s1. eReference A. 77 cases the limiting chemical shifts and the stability of the complex are not affected either by a change in the salt concentration or by a change in the anion. It is evident, therefore, that at low concentrations used in this study, the formation of the complex is unaffected by ion pairing. We can assume that in solvents of higher donicities and/or higher dielectric constants the same situation will exist. The large divergence in DB2HCB'Cs+ complex limiting chemical shifts (a range of about N6 ppm) probably indi- cates that the cesium ion is not insulated from the solvent by the ligand. The complex formation constant in methanol solution is in satisfactory agreement with the previously reported value (A). 3.3.3. DB24C8 Complexes with T1+ The thallium nucleus, 205Tl, has very favorable prop- erties for NMR studies. It has a spin of I = 1/2 and its chemical shift is very sensitive to small changes in the chemical environment. The solvent dependence of the thallium-205 chemical shift is over 2600 ppm (116,120) in comparison to a chemical shift range of about 8 ppm for 7L1 (123,12u), 30 ppm for 23Na (123,121), and 130 ppm for 133Cs (125,126). The thallous ion is a useful probe because its chemistry is very similar to that of the alkali ions (127). In particular, the chemical properties and ionic radii of Tl+ and K+ ions are very close (1.5“ X and 78 1.11 1, respectively) (128) so that thallium(I) ion can be used as a NMR probe for potassium. Because of these interesting properties, we were interested in studying the complexation of thallous ion with dibenzo-2h-crown-8 in various nonaqueous solvents. The thallium-205 chemical shift was measured as a function of DB2HC8/Tl+ mole ratio in nitromethane, aceto- nitrile, acetone, methanol, DMF, DMSO, and pyridine solu- tions at 30°C. The thallous ion concentration was main- tained constant at 0.005 M_in all cases. The mole ratio data are given in Table 1A and the variations of the 20511 chemical shift as a function of the ligand/Tl+ mole ratio are shown in Figure 10. In poor solvating solvents such as nitromethane, acetonitrile, and acetone, the addition of the ligand to the thallous ion solution causes a quite linear change in the 20511 chemical shift (upfield or downfield) until a mole ratio of l is reached. After the mole ratio of 1, further addition of the ligand does not affect the thallium-205 resonance. This behavior indicates the formation of a very stable l:l complex between the thallium(I) ion and the ligand in the above solvents. 'On the other hand, in methanol solution the mole ratio plot shows some curvature around the mole ratio of l and a limiting value is obtained at mole ratios greater than 2.5, indicating formation of a weaker 1:1 complex than that in the previous solvents. In solvents of high 79 Table 11. MOle Ratio Study of Dibenzo-21-Crown-8 Complexes with O-OO5E.T10101 in Various Solvents at 30°C. NM AN Ac MeOH L/Tl+ 6 + 6 + 6 + (ppm) L/Tl (ppm) L/Tl (ppm) L/Tl (ppm) 0.00 366.6 0.00 209.8 0 00 206.1 0.00 -lo7.5' 0.73 281.9 0.31 222.1 0 36 226.1 0.16 -7.7 0.97 250.1 0.70 237.8 0 72 212.8 0.75 60.7 1.11 216.9 0.85 213.1 0 82 219.9 0.86 78.6 1.22 216.9 0.96 215.6 0 97 253.1 1.06 118.2 1.65 216.9 1.09 219.1 1 10 257.0 1.15 131.8 1.88 216.7 1.21 219.8 1.31 258.3 1.28 113.1 2.15 216.9 1.52 219.8 1.56 258.9 1.59 166.3 2.68 216.9 1.88 219.9 1.87 259.7 1.79 176.7 2.19 219.9 2.23 259.6 2.06 187.6 2.72 250.0 2.80 260.1 2.70 192.3 DMF DMSO Py + + + L/Tl 6(ppm) L/Tl 5(ppm) L/Tl 6(ppm) 0.00 -l90.2 0.00 -319.3 0.00 651.1 0.35 -l78.3 0.15 -320.7 0.38 603.1 0.78 -163.5 0.80 -321.7 0.67 569.1 0.91 -156.3 0.87 -322.0 0.87 511.9 1.12 -151.5 0.98 -322.3 0.99 529.3 1.22 -118.0 1.10 -322.7 1.08 516.7 1.16 -139.3 1.29 -323.1 1.22 198.8 1.80 -126.2 1.61 -321.2 1.59 161.1 2.08 -115.2 1.83 -321.8 1.81 111.5 2.19 -113.6 2.37 -325.7 2.35 398.5 2.77 -102.1 2.81 -326.6 2.80 360.2 80 600 400 NM 200 MeOH 0.— DMF -20 l I I I I ((1.0 0.5 1.0 1.5 2.0 is 3.0 [DB24C81/[TF] Figure 10. Thallium-205 Chemical Shifts XE: [DB21081/[T1+] Mole Ratio in Different Solvents. 81 donicities, such as DMF, DMSO, and pyridine, upon addition of the ligand there is a gradual shift of the 205T1 reson- ance which does not reach a limiting value even at a mole ratio of 3 (Figure 10), indicating formation of a weak complex in these solvents. The formation constants and the limiting chemical shifts of the l:l complex of the ligand with the thallous ion in various solvents are given in Table 15. It is immediately obvious that the stability of the complex is very much dependent on the nature of the solvent. The stability of the complex increases with decreasing Gutmann donor number of the solvents. It is interesting to note that in con- trast to the sodium and cesium ion complexes with DB21C8, in pyridine solution the DB21C80T1+ complex is expectedly weak. The thallous ion as a "soft acid" can strongly interact with the nitrogen atom of pyridine which is a "soft base" so that the resulting DB2HC8oTl+ in this solvent is weak. As shown in Table 15, the formation constants of the complex in acetonitrile and methanol solutions are in a good agreement with the values reported by Hofmanova gt él' (35) who used polarographic technique for the measurements. The limiting chemical shifts of the complexed thallous ion in solvents of low and medium donicities such as nitromethane, acetonitrile, acetone, and methanol are close to each other (Figure 10). This behavior indicates 82 Table 15. Formation Constants and the Limiting Chemical Shifts of DB21C8-Tl+ Complexes in Various Solvents. Solvent Log Kf 611m(ppm) Nitromethane >5 215.00 Acetonitrile 1.81:0.05 250.110.l Acetonitrile 1.80a ---- Acetone - 1.15:0.05 260.510.l Methanol 3.19:0.07 215.2:3.8 Methanol 3.10a ---- Dimethylformamide 1.1610.2l 336 :91 Dimethylsulfoxide 082108-Cs+ > DB2108-Na+. For large crown ethers such as dibenzo-30-crown-10 and dibenzo—21-crown-8 which are capable of formation of three-dimensional "wrap around" complexes with cations (33), we expect the size of the cation to play an important role in the complexation reaction. The complexation of a cation 81 Table 16. Formation Constants of 1:1 Complexes of Na+, 03+, and Tl+ Ions with Dibenzo-21-Crown-8 in Various Solvents. Log Kf Solvent Cs+ Tl+ Na+ Nitromethane 1.11:0.08 >5.0 3.7110.12 Acetonitrile 3.91:0.07 U.8110.05 2.95:0.07 Acetone 3.71:0.09 U.lSi0.0S ---- Methanol 3.60:0.05 3.1910.05 ---- Dimethylformamide 2.10:0.01 1.1610.2l 10.0 Dimethylsulfoxide 1.61:0.01 .m m=.o em :m.m m=.o we mm.HH H=.o ow om.m oo.o ma sa.s oo.o ma mm.mH oo.o Anmvmxaoo Agoove +mzxd Anmvmxaoo Asodve +mzxd Anmvmxaoo “shove +m2\q ca za :2 .000m um mpcm>aom mSOfipm> ca sammmz_ammo.o spas moxoaosoo euszoaonamnonsooao co sosom oaomm ofloz .sH oHome 88 mmm 00.0 00.m 00m 00.0 0m.m 00H 0H.0 00.0 00 00.0 mH.m mam mm.0 m0.H 03H 0H.0 00.H mm 00.0 00.H 000 00.5 mm.H 00H 01.0 mm.H 00 H0.: 0m.H 0:0 0:.» 0H.H NHH 0H.0 mm.fi as m0.0 0m.H 000 00.0 00.0 00 0H.0 ~0.H 00 00.0 m0.H 0H0 as.m 00.0 mm mH.0 00.0 H0 00.: 00.0 00H 00.: 00.0 mm 00.0 00.0 mm 00.0 H0.0 maa m=.m mm.0 m0 m0.0 0m.0 0m 00.: 00.0 0H 0H.Hu 00.0 m: H0.0- 00.0 00 00.: 00.0 Asavm\flea Aecaos +12\a AN00N\H90 Aeaavc +12\a A600~\H>< Aecave +QZ\4 00 0020 0:0 .0cssfieoo0 .NH magma 89 6- o DMF 4.. 2.. 0' DMSO -2 1 1 J, I I I 00 0-5 1.0 1.5 2.0 2.5 3-0 [0821C 7]/[No+] Figure 11. Sodium-23 Chemical Shifts E- [DB2lC7]/[Na+] Mole Ratio in Different Solvents. 9O abilities which can strongly compete with the ligand for the cation. 0n the other hand, a considerable change in 23Na chemical shifts was observed in nitromethane, acetonitrile, acetone, and pyridine solutions upon the addition of the ligand indicating a change in the chemical environment of the nucleus. The diamagnetic shift of the sodium-23 resonance in acetonitrile, acetone, and pyridine solution (paramagnetic in nitromethane) begins to level off at a mole ratio of about 1, indicating formation of a 1:1 complex between the ligand and the sodium ion. In all cases, the line width of the 23Na resonance at the half height (Avl/Z) increases almost linearly with the ligand/Na+ mole ratio until a mole ratio of about 1 is reached, and then it remains almost constant upon further addition of the ligand (Table 17). This is not surprising because for a nucleus such as sodium with an appreciable quadrupole moment we expect the line width of the resonance signal to be sensitive to the electrical field gradient around the nucleus. The solvated sodium ion has a symmetrical environ- ment and, therefore, the line width of the resonance signal is small (about 10 to 20 Hz depending on the solvent used). Complexation of the sodium ion with the ligand creates an unsymmetrical environment around the cation which results in the broadening of the resonance signal. As the DB2IC7/Na+ mole ratio increases, the amount of the complexed sodium 91 23Na resonance signal ion increases and, therefore, the becomes broader. After a mole ratio of l is reached, essentially all of the cation is complexed and further addition of ligand has no effect on the sodium-23 resonance. This behavior, once more, confirms the formation of a stable 1:1 complex between DB2lC7 and sodium ion. The formation constant and the limiting chemical shifts of DB21C7-Na+ complex in various solvents at 30°C are shown in Table l8.’ With the exception of pyridine there is an inverse relationship between the donicity of the solvent and the stability of the complex. It has been pointed out previously, however, that being a soft nitrogen donor, pyridine does not strongly solvate a hard ion such as sodium ion (80) and, therefore, a stable complex can be formed between the ligand and the sodium ion in this solvent. The limiting chemical shifts of the complexed sodium ion in nitromethane, acetonitrile, acetone, and pyridine solutions are very close, indicating that the sodium ion is enclosed inside the ligand's cavity so that the solvent molecules can barely interact with the cation. 3.1.2. DB21C7 Complexes with Cs+ Cesium-133 NMR was used to study the complexation of the cesium ion with dibenzo-2l-crown-7 in various non- aqueous solvents. The 133Cs chemical shift was determined :15 a function of the ligand to cesium ion mole ratio. 92 Table 18. Formation Constants and the Limiting Chemical Shifts of DB2lC7-Na+ Complexes in Various Solvents Solvent Log Kf 611m(ppm) Nitromethane 3.11:0.05 9.35:0.01 Acetonitrile 2.7810.08 10.2910.02 Acetone 2.2810.08 9.88:0.06 Dimethylformamide- ~0.0 ---- Dimethylsulfoxide ~0.0 ---- Pyridine 2.56:0.05 9.1710.06 93 The results are shown in Table 19 and in Figure 12. The cesium ion concentration was maintained at 0.005 M, Upon addition of the ligand, the line width observed shows a slight increase from about 1 Hz to 11 Hz. The complexation was studied up to a mole ratio of about 3, where a limiting value for the chemical shift of the complexed cesium ion is reached in nitromethane, acetonitrile, acetone, methanol, and pyridine solutions. The variations of the cesium-133 chemical shift as a function of DB2lC7/Cs+ mole ratio in the above mentioned solvents show a single inflection point at a mole ratio of about 1, indicating formation of a 1:1 complex between the ligand and the cesium ion. No evidence for formation of a 2:1 (ligand to metal ion) complex was observed in any of the solvents used. While the existence of a 2:1 complex of the ligand with cesium ion in the solid crystalline state has been demon- strated (23), it does not follow necessarily that it also exists in solution. Frensdorff (1), however, has reported formation of a very weak 2:1 complex between DB2lC7 and cesium ion in methanol solution. The formation constant and the limiting chemical shifts obtained from the computer analysis of the mole ratio data in various solvents are shown in Table 20. The large difference in the limiting chemical shifts of the complex in different solvents is a good indication that the metal ion remains exposed to the solvent. The cesium ion is 91 Table 19. Mole Ratio Study of Dibenzo-21-Crown-7 Complexes with 0.005! CsSCN in Various Solvents at 30°C. NM AN AC MeOH + + + + L/Cs 6(ppm) L/Cs 6(ppm) L/Cs 6(ppm) L/Cs 6(ppm) 0.00 57.89 0.00 -3l.93 0.00 19.18 0.00 16.97 0.26 50.13 0.35 -21.79 0.30 16.16 0.30 39.53 0.50 13.37 0.61 -l9.06 0.56 13.82 0.51 32.86 0.77 35.22 0.82 -15.19 0.79 11.65 0.68 29.92 0.91 30.75 0.92 —l3.17 0.89 10.72 0.90 21.61 1.11 27.37 1.01 -ll.76 1.03 9.67 1.03 22.10 1.28 25.11 1.22 -9.U7 1.12 8.9“ 1.1“ 21.55 1.18 21.86 1.33 -9.36 1.20 8.78 1.28 20.62 1.80 21.21 1.50 -9.05 1.13 8.12 1.16 19.90 2.25 21.15 1.83 -8.7N 1.90 8.2“ 1.85 19.66 2.67 21.07 2.52 -8.52 2.37 88.20 2.12 19.53 3.21 -8.10 3.16 8.10 3.01 19.31 DMF DMSO Pya + + + L/CS 6(ppm) L/CS 6(ppm) L/CS 6(ppm) 0.00 2.51 0.00 -66.59 0.00 21.53 0.37 2.20 0.30 ~63.18 0.17 19.80 0.69 1.88 0.72 -58.19 0.50 11.12 0.86 1.61 0.93 -55.89 0.77 3.91 0.99 1.50 1.06 -51.27 0.91 -0.67 1.19 1.91 1.20 -53.25 0.99 -2.30 1.31 1.36 1.90 -50.92 1.1“ -5.U7 1.95 1.27 1.82 -17.36 1.29 -5.63 2.35 1.19 2.37 -13.72 1.51 -5.78 2.9“ 1.11 3.06 -39.73 1.88 —5.82 2.37 -5.91 2.97 -6.05 60 95 NM MeOHI f 4 . a —.> AIPY AN’“ J J L 0.0 0.5 1.0 1.5 2:6 285 380 [DB21C7]/[Cs"] Figure 12. Cesium-133 Chemical Shifts vs. [DB2107]/[Cs+] Mole Ratio in Different SolVEhts. 96 Table 20. Formation Constants and the Limiting Chemical Shifts of DB2107°Cs+ Complexes in Various Solvents. Solvent Log Kf 611m(ppm) Nitromethane 1.11i0.07 23.78i0.05 Acetonitrile 3.95:0.01 -8.1720.02 Acetone 3.93:0.06 7.98:0.02 Methanol 3.96:0.06 19.07:0.01 Methanol 1.20a ---- Dimethylformamide 2.81:0.10 0.92:0.01 Dimethylsulfoxide 1.72:0.01 -0.96:3.18 Pyridine 1.27:0.07 -6.2110.03 aReference 1. 97 known as a large cation with a low charge density which cannot be strongly solvated by the solvent molecules. Thus it is not surprising if the stability of the complex in nitromethane, acetonitrile, acetone, methanol, and pyridine is only marginally dependent on the nature of the solvent. In DMF and DMSO, however, the DB2lC-Cs+ complex is much weaker than that in the above solvents. Dimethylformamide and dimethylsulfoxide are solvents of high donicities, and their interactions even with a large cation such as cesium ion is still strong enough to compete with the ligand for this cation and to prevent formation of a strong complex. The value obtained for the stability constant of the DB2lc7-Cs+ complex in methanol solution at 30°C is in a satisfactory agreement with the literature reported value in the same solution at 25°C (1) (Table 20). 3.1.3. DB2107 Complexes with T1+ In order to determine the stoichiometry and the stability of the thallium(I) complex with dibenzo-2l- crown-7 in various solvents, the thallium-205 NMR chemical shift was measured as a function of DB21C7/Tl+ mole ratio. The concentration of thallium(I) perchlorate was maintained at 0.005 M, while the concentration of the ligand was varied from zero to about 0.015 M. The data are given in Table 21 and the mole ratio plots are shown in Figure 13. In nitromethane, acetonitrile, and acetone solutions, the 98 Table 21. Mole Ratio Study of Dibenzo-21-Crown—7 Complexes with 0.005M’T1C10u in Various Solvents at 30°C. NM AN Ac + + + L/Tl 6(ppm) L/Tl 6(ppm) L/Tl 6(ppm) 0.00 319.6 0.00 213.9 0.00 150.3 0.71 319.2 0.37 231.1 0.12 202.8 0.86 315.1 0.85 260.5 0.51 216.8 1.01 310.1 1.05 272.2 0.77 216.5 1.11 309.0 1.15 271.1 0.91 262.1 1.28 308.8 lL32 271.3 1.06 271.6 1.13 308.5 1.50 271.6 1.21 275.0 1.88 308.1 1.98 271.7 1.51 277.3 2.37 308.3 2.55 271.6 1.63 277.9 3.11 308.1 3.21 271.5 2.03 278.1 2.57 278.2 3.09 278.3 MeOH DMF DMSO + + + L/Tl 6(ppm) L/Tl 6(ppm) L/Tl 6(ppm) 0.00 -112,o 0,00 -l18.5 0.00 -212.9 0.28 -39.u 0.10 -9u.3 0.27 -218.3 0.52 32.0 0.70 -55.1 0.51 -223.1 0.76 103.7 0.86 -36.0 0.82 —230.0 0.87 130.0 1.03 -15.2 0.92 -231.9 0.96 151.1 1.13 -6.7 1.01 -233.1 1.10 182.9 1.33 15.3 1.16 -236.2 1.23 188.0 1.85 11.1 1.38 -210.3 1.18 200.2 2.12 75.5 1.60 -216.0 1.75 201.1 3.11 101.1 1.80 -218.5 2.20 207.5 2.17 -261.0 2.90 209.1 3.09 -273.3 99 N1 3mJ- .Ac 1 - 7:.- aMeOH zoo * DMF mo - 8 (PM)! 0 .- -]00 .— -2oo DMSO 1 l L l 1 L 00 05 10 LS 10 25 10 [Dismal/[n+1 Figure 13. Thallium—205 Chemical Shifts gs. [DBZlC7]/[T1+] Mole Ratio in Different Solvents. 100 thallous ion resonance shifts either upfield or downfield as the concentration of the ligand increased. Once again mole ratio plots show a break at a mole ratio of 1, indi- cating formation of a 1:1 complex. The break is more pro- nounced in the nitromethane and acetonitrile curves compare to that in acetone and methanol plots which is indicative of formation of a stronger complex in the former solvents. No other break is found, indicating that if any 2:1 (ligand to metal ion) complex formation occurs it is negligible. In solvents of high donicities such as DMF and DMSO only a gradual shift of thallium-205 resonance is observed upon addition of the ligand which does not reach a limit- ing value even at a mole ratio of about 3. This behavior indicates formation of a weak complex between the ligand and the thallous ion in these solvents. The formation constant and the limiting chemical shift of the 032107-11+ complex in various solvents are given in Table 22. It is obvious that the nature of the solvent has a large effect on the stability of the complex. There is an inverse relationship between the Gutmann donor number of the solvent and the complex formation constant. Except for the acetonitrile and acetone cases the complex limiting chemical shifts are different in different solvents, indicating that the complexed thallous ion remains exposed to the solvent. For stable complexes with Kf > 105, the 205T1 Chemical shift-mole ratio plots consist of two straight lines 101 Table 22. Formation Constants and the Limiting Chemical Shifts of DB2lC7-T1+ Complexes in Various Solvents. Solvent Log Kf 611m(ppm) Nitromethane >5 308.1 Acetonitrile >5 271.5 Acetone 1.71:0.08 278.5i0.1 Methanol - 3.97:0.03 212.710.3 Dimethylformamide 2.18:0.02 233.8:8.1 Dimethylsulfoxide 0.63:0.15 ---- 102 intersecting at 1:1 mole ratio. These kinds of plots cannot be analyzed by our technique, and in such cases we can only conclude that log Kf > 5. This behavior is observed for the DB2107-Tl+ complex in nitromethane and acetonitrile solutions which is not surprising since both nitromethane and acetonitrile are poor solvating solvents with respective Gutmann donor number of 2.7 and 11.1. 3.1.1. Conclusions The formation constants of the 1:1 complexes of Na+, 03+, and Tl+ ions in various solvents are compared in Table 23. In solvents of low and medium solvating abilities such as nitromethane, acetonitrile, acetone and methanol, the stabilities of DB21C7 complexes with sodium, cesium, and thallium(I) ions decrease in the order DB21C7-T1+ > DB21C7-05+ > DB21C7-Na+. The ionic radii of the above cations vary in the order Na+ < T1+ < Cs+. Cesium ion has a diameter of 3.68 1 (128) which is Just the right size to fit conveniently inside the cavity of dibenzo-2l- crown-7 with the size of 3.1-1.3 K (1), while sodium ion (diameter 2.21 3) is too small for the ligand's "hole". Thus the DB2lC7-Cs+ complex is more stable than the DB21C7-Na+ complex. In the case of DB2lC7-Cs+ complex the crown ether only occupies the equatorial coordination sites of the cation. Thus the complexed cesium ion remains exposed to the solvent molecules from the axial positions 103 Table 23. Formation Constants of 1:1 Complexes of Na+, 03+, and T1+ Ions with 032107 in Various Solvents. Log Kf Solvent Na+ Cs+ 11* Nitromethane 3.1110.05 1.11:0.07 >5 Acetonitrile 2.78:0.08 3.95:0.01 >5 Acetone 2.28:0.08 3.93:0.06 1.71:0.08 Methanol ---- 3.96:0.06 3.97:0.03 Dimethylformamide ~0.0 2.81iO.10 2.18:0.02 Dimethylsulfoxide ~0.0 1.72:0.01 0.63:0.15 Pyridine 2.5610.05 1.27:0.07 ---— 101 and, therefore, the complex limiting chemical shift is strongly solvent dependent (Figure 12). In the case of DBZlC7-Na+ complex, however, because of the small size of the cation, the ligand is able to twist around the sodium ion to form a three—dimensional array and thus insulate it from the solvent molecules. Conse- quently, the limiting chemical shift of the complexed cation is essentially solvent independent (Table 18). In the case of the thallium(I) complex, despite the more inconvenient relative sizes of the cation and the ligand than the cesium ion case, the DBZlC7-Tl+ complex is more stable than the DB210-Cs+. This is not surprising since the thallous ion is known to bond to the oxygen atoms of the macrocyclic ligands by an ion-dipole interaction with a covalent contribution (129) which results in the forma- tion of a very strong complex between the cation and the ligand. In solvents of strong solvating abilities such as DMF and DMSO no evidence for the formation of a complex between the ligand and the sodium ion was observed. The cesium and thallous ions, however, form weak complexes with the ligand in these solvents, but in this case the DB21C7-CS+ complex is more stable than the DB21C7-Tl+ complex. The smaller stability constant of the thallous ion complex inay be related to the preference of thallous ion as a soft acid for the nitrogen and sulfur atoms as soft bases 105 in the DMF and DMSO structures respectively. Pyridine is an exceptional case where despite the high donicity (Gut- mann donor number of 33.1) of the solvent, the resulting complexes of the ligand with sodium and cesium ions are unexpectedly stable. The reason for this behavior has been discussed previously. 3.5. DISCUSSION The stability constants of Na+, 08+, and Tl+ ion com- plexes with dibenzo-2l-crown-7, dibenzo-21-crown-8, and dibenzo-30—crown-10 in various solvents at 30°C are shown in Table 21. It is immediately obvious that in all cases the nature of the solvent has a great effect on the stabilities of the complexes. The magnitude of the solvent effect on the complex formation constants decreases in the 4. order of Tl+ > Na > Cs+. Cesium ion is a large cation with a low charge density which cannot strongly interact with the solvent and, therefore, the stabilities of the cesium ion complexes are only marginally dependent on the nature of the solvent. In the case of the sodium ion complexes the solvent effect is more pronounced, which can be related to the smaller Size and consequently to the higher charge density of the sodium ion than that of the cesium ion. Since the thallium(I) ion is bonded to the solvent Inolecules by an ion-dipole interaction with a covalent 106 :o.ou:m.a IIII oa.ouaa.: mo.onoo.= No.0wpm.: oa.onmm.m mo.o«mm.m maficfipmm oo.Hv mfi.ommm.o IIII :o.o«am.a zo.ommw.a 0.08 0.08 omza Hm.o«mH.H mo.o«ma.m IIII :o.owoa.m oa.ow=m.m 0.08 0.08 min mo.owma.m mo.ou~m.m wo.owma.z mo.owom.m mo.owmm.m IIII IIII Hocmnpmz mo.onma.: mo.oaan.= no.0Hmm.m mo.ona>.m mo.oumm.m IIII mo.osmm.m ocopoo« mo.onam.: mA mo.owmm.m no.0«zm.m no.owmm.m uo.o«mm.m mo.o«mu.m mafippacoumo< omA mA mo.ouom.n wo.owaa.= No.ouza.: ma.ou:w.m mo.ouza.m ocmnpothpdz mozmmo scamma oaoommo mozmmn noammo mozmma woamma pco>aom +HB +mo m2 ox moo .ooom pm mucm>aom msofihm> ca oHoome cam .mozmma .woamma spa: mCOH +HB new .+mo .+mz mo moonQEoo Hua no mpcmumcoo coapmapom .zm magma 107 contribution, it is not surprising that the stabilities of the complexes depend strongly on the nature of the sol- vent. With the exception of the sodium and cesium ion com- plexes in pyridine solution, there is an inverse relation- ship between the donicity of the solvent and the stability of the complex. In the case of pyridine, however, we have a solvent with the highest donor strength (142;, Gutmann donor number of 33.1) and yet the sodium and cesium ion complexes are strong in this medium. It has been pointed out previously that pyridine as a "soft" nitrogen donor does not solvate strongly "hard" cations such as an alkali ion (80,130). As can be seen, sodium ion forms a more stable complex with DBZ1C8 than with DB21C7. While the cavity sizes of both ligands are too large for the small sodium ion, the increased number of the oxygen atoms, as binding sites, in DBZ1C8 seems to play an important role in the complexa- tion. As discussed in previous sections (3.3.1 and 3.1.1) the limiting chemical shifts of the complexed cations are relatively solvent independent in both cases indicating that the sodium ion is insulated from the solvent by the jligands. According to this picture it is evident that iDBZ1C8 with more binding sites can form a more stable Icomplex with the sodium ion than DB21C7. In the case of the thallous ion complexes the opposite behavior is 108 seen: 032107-11+ is more stable than DB21C8°T1+. In this case the relative sizes of the ligand to the cation seem to be the key factor in the complexation reactions. The thallous ion with the size of 3.08 K has a more convenient fit inside the cavity of 002107 (with the size of 3.1—1.3 1) than with DB21C8 which has a cavity size greater than 1 X. With the exception of the DB3OClO-Cs+ complex in aceto- nitrile solution the stabilities of the cesium ion complexes with the ligands decreases in the order DB3OC10-CS+ > + in all solvents used. Dibenzo-30- 002107-Cs+ > DB21C8-Cs crown-10 is a large molecule with enough oxygen atoms in the ring which can form a stable "wrap around" complex with cesium ion (Sec. 3.2.2). The Sizes of the dibenzo-21- crown-7 cavity and the cesium ion are very close so that the cation can be held by the ligand to form a stable complex, which is expectedly weaker than DB3OC10-Cs+ complex because of its two-dimensional structure. Dibenzo- 21—crown-8 has neither a long enough chain to form a three- dimensional complex with the cesium ion nor a convenient cavity size to hold the cation as tight as dibenzo-21- crown-7 can. Thus the resulting complex is the weakest in the series. CHAPTER 1 CESIUM-133 NMR STUDY OF THE THERMODYNAMICS OF THE COMPLEXATION 0F DIBENZO-30-CROWN-10, DIBENZO-21-CROWN-8, AND DIBENZO-21-CROWN-7 WITH CESIUM ION IN NONAQUEOUS SOLVENTS 109 1.1. INTRODUCTION In order to have a deeper understanding of the thermo- dynamics of the complexation reactions, it is useful to consider separately the enthalpic and the entropic contribu- tions to the reaction. The change in free energy of complexa- tion, AG°, can be divided into two parts: the change in enthalpy, AH°, and the change in entropy, A80, of the reaction: AG° = AH° - TAS° (l) The complex can be stabilized by meeting one of the follow- ing requirements: (a) AH° < 0 and dominant, TAS° < 0 (enthalpy stabilized); (b) AH° > 0, TAS° > 0 and dominant (entropy stabilized); and (0) either AH° < 0 and dominant, TAS° > 0 or AH° < 0, TAS° > 0 and dominant (both enthalpy and entropy stabilized). Presently available thermodynamic data for the com- plexation of alkali metal ions with uncharged ligands such as crowns and cryptands are not detailed enough to give a clear picture of the thermodynamic behavior of the com- plexes. The results of the thermodynamic studies, reported so far, Show that most of the alkali ion complexes with crowns and cryptands (18) are enthalpy stabilized but entropy destabilized. In particular, the reported 110 111 thermodynamic data for the complexation of the cations with large crown ethers (112;: 2l-crown-7 and larger) are quite sparse. To the best of our knowledge, there are Just four papers available (96,33,31,l3l), reporting the thermodynamic parameters fortflmecomplexation of large crowns. In previous work, Lehn and coworkers (132) and Cahen _§|al. (110) both noted that the observed chemical shift of the complexed metal ion varied with temperature. Cesium- 133 and lithium-7 NMR were used by E. Mei (112) and by A. Hourdakis (133) to determine the thermodynamic parameters for the complexation of the cesium and lithium ions with some crowns and cryptands in nonaqueous solutions. The results of the study of the stoichiometries and +, Cs+, and T1+ ion complexes with stabilities of Na+, K dibenzo-30-crown-10, dibenzo-21-crown-8, and dibenzo-2l- crown-7 in various solvents were discussed in chapter 3. In this chapter we will attempt to evaluate the thermo- dynamic parameters of the complexation reaction between cesium ion and the above mentioned crown ethers. 1.2. DB3OC10 COMPLEXES WITH Cs+ The method for determining of the thermodynamic values ‘was based on the temperature dependence of the complex ‘formation constant. The complex formation constant is 112 related to the relevant thermodynamic parameters by the following relationships: AGO = AHO - TASO (l) o o in Kf = - %gr-+ A: (3) Thus a plot of in Kf vs. l/T gives a straight line with a AH° AS° o slopt of - R and an intercept of —R—, provided that AH is independent of temperature over the temperature range considered. The variation of the cesium-133 chemical shift was measured as a function of DB3OClO/Cs+ mole ratio in nitro- methane, acetonitrile, acetone, methanol, and pyridine at different temperatures. In all cases studied, only one population average signal was observed indicating that the exchange of the metal ion between the two sites (112;: free ion in the bulk solution and the complexed ion) is faster than the NMR time scale. The data are given in Table 25 and the mole ratio plots at different tempera- tures are shown in Figures 11-18. In all five solvents used, i;g;, nitromethane, aceto- nitrile, acetone, methanol, and pyridine, the curvature in the mole ratio plots increases with increasing the temperature. This behavior indicates the existence of 113 Table 25. Cesium-133 Chemical Shifts of 0.005M_Cs Ion in the Presence of DB3OC10 at Various Tempera- tures. Solvent: Nitromethane DB3QClO/Cs+ Temperature, °C 10 30 15 60 70 0.00 58.21 59.71 61.71 63.01 61.81 66.51 0.23 18.61 50.61 . 52.11 53.81 55.91 57.21 0.50 37.71 38.21 39.72 11.30 13.21 15.12 0.71 27.10 28.51 30.61 32.81 31.80 36.11 0.81 23.31 21.22 25.70 27.91 30.72 32.71 0.96 19.82 21.11 22.00 21.22 27.00 29.62 1.15 17.70 19.12 20.11 21.80 21.81 26.10 1 30 17.11 18.82 19.10 21.22 23.71 21.71 1.19 17.11 18.30 19.10 20.21 22.60 23.52 1.68 17.11 18.22 19.11 20.21 22.00 22.11 1.80 17.11 18.30 19.00 19.82 21.81 22.20 2.08 17.12 18.22 18.92 19.61 21.30 22.11 Solvent: Acetonitrile 0030010/Cs+ Temperature, °C 18 30 15 60 77 0.00 -35.10 -33.65 -32.11 -30.58 -26.00 0.25 -26.75 -25.21 -23.10 -22.33 —l9.00 0.50 -17.13 -16.20 -l1.73 -l1.01 ~12.33 0.59 -l1.81 -13.57 -12.71 -11.39 -9.85 0.70 -10.62 -9.77 -9.07 -8.36 -7.91 0.81 -6097 -6028 -60u3 -5082 -507“ 0.98 -1.08 -0.85 -1.70 -1.62 -2.25 1.12 3.11 2.79 2.02 1.98 -0.10 1.19 1.12 1.12 3.26 2.18 1.10 1.38 7.37 6.98 6.13 5.27 1.19 1.60 8.61 8.51 8.00 7.29 6.05 2.00 9.55 9.55 9.71 9.28 8.02 111 Temperature, °C 20 30 50 29.13 30.87 33.16 25.78 26.87 29.73 21.52 22.38 21.77 20.20 20.97 23.38 18.19 19.66 22.13 17.31 18.26 21.06 17.03 17.87 20.58 16.10 16.87 19.53 15.70 16.18 18.80 15.65 16.11 18.26 15.63 16.25 18.03 Temperature, °C 30 15 60 Table 25. Continued. Solvent: Acetone 0030010/Cs+ -10 10 0.00 25.13 28.11 0.30 21.91 21.16 0.61 17.87 20.20_ 0.75 16.55 18.57 0.88 15.31 17.26 0.99 13.72 16.17 1.07 13.26 15.86 1.23 13.01 15.21 1.36 12.91 11.93 1.53 12.91 11.93 1.86 12.91 11.93 Solvent: Methanol 0030010/Cs+ 0 15 0.00 12.10 11.37 0.22 35.05 37.69 0.56 25.05 27.38 0.80 17.69 20.71 0.86 17.00 19.97 0.98 13.97 17.30 1.07 13.19 16.25 1.26 13.15 15.11 1.37 13.11 15.28 1.60 13.11 15.15 1.71 13.11 15.08 16.00 17.23 50.25 39.75 11.18 15.38 30.02 32.26 37.16 23.71 27.30 32.11 23.31 26.25 31.25 20.17 21.71 28.95 19.15 23.85 27.99 18.26 21.50 26.15 17.61 20.57 25.67 17.11 19.93 21.13 17.11 19.75 23.97 115 Table 25. Continued. Solvent: Pyridine 0030010/Cs+ Temperature, °C 0 10 3O 50 65 85 0.00 27.17 30.05 31.39 38.12 11.68 11.21 0.22 23.69 25.50 28.57 31.91 31.70 37.99 0.17 18.65 20.28 _ 23.77 26.09 27.95 30.82 0.66 11.61 16.56 18.57 20.82 22.52 25.17 0.80 11.99 12.60 11.77 16.18 17.71 21.59 0.86 11.81 11.82 13.69 15.50 16.86 20.13 1.03 9.02 9.80 11.11 13.06 11.16 17.25 1.10 9.25 9.73 11.20 12.76 11.32 17.22 1.27 9.19 9.72 10.90 11.83 13.22 16.01 1.15 9.17 9.62 10.66 11.75 12.81 15.38 1.77 9.19 1.61 10.51 11.11 12.28 11.78 2.18 9.18 9.59 10.51 11.10 12.15 11.37 116 NITROMETHANE 60k 50”- E‘ o. 9: co 40- 30- 700 so. 20 I- 450 it" L l 1 L 1 l 1 L 1— 1 1 013(12 CMQ O£5<18 I!) L2. h4 Iii L8 2£>12 [ DB3OC10]/[Cs*] Figure 11. Cesium-133 Chemical Shifts vs. [DB3OClO]/[Cs+] Mole Ratio in Nitromethane at Different Tempera- ures. 117 ACETONITRILE u; m» f, £0 502 77 i, 7.2..) / l l 0.0 0.5 Ito Ls 2.0 [0830001161 Figure 15. Cesium-133 Chemical Shifts vs. [DB3OC10]/[Cs+] Mole Ratio in Acetonitrile EE'Different Tem- peratures. ACETONE 1 30 l 1 1 6 (.0me 20 '- 50° 30° 20° - 4 : —: 10: i t -10 10 '- l l l i as 1.0 1.5 20 [0830001165 0.0 Chemical Shifts v_s_. [DB3OClO]/[Cs+] Mole Ratio in Acetone at Different Temperatures. Figure 16. Cesium-133 119 5O METHANOL 40 I E‘ a. 2: 6° 30*- 60' 20" 45‘ 30‘ ISP 0. IO- L l L J l I l l I CLO 022(L4I(16 CHE Ix) L2 64 M5 L8 [08300011116] Figure 17. Cesium-133 Chemical Shifts 1g. [DBBOClO]/[Cs+] Ratio in Methanol at Different Temperatures. 120 PYRIDINE 4o1~ 30 O E: a. m 20-— as- 659 so- 30- IC-- 197 l l l l 1 l L l L l J 0.0 0.2 0.4 0.6 0.8 to L2 L4 l.6 L8 2.0 2.2 [moon/[cg] Figure 18. Cesium-133 Chemical Shifts vs. [DB3OClO]/[Cs+] Mole Ratio in Pyridine at DIITerent Temperatures. 121 an exothermic reaction between the cesium ion and the ligand. The formation constants were computed by the KINFIT program as described previously and the results are listed in Table 26. It is readily seen that lowering the temperature in- creases the stability of the complex. In most cases at lower temperatures the formation constants were greater than 105 and their precise values could not be determined by our technique. Plots of in Kf vs. l/T for the five systems are shown in Figure 19. The slopes and intercepts of the straight lines are calculated by linear least squares fitting of the data, and the corresponding thermodynamic parameters are listed in Table 27. It is seen that the results ob- tained in methanol solutions agree reasonably well with the results of Chock (32). It is interesting to note that while the stability (or AG° values) of the complex is not very sensitive to the solvent (at least in the case of the five solvents studied here), the enthalpy and the entropy values vary Significantly with the solvent. Since the cesium ion is rather weakly solvated because of the low charge density of the cation, it is not surprising that the free energy, AG°, of the complexation is only marginally dependent on the nature of the solvent. However, the solvation of the ligand would be different in different solvents and because of that the enthalpy of the complexa- tion is solvent dependent. The entropy of the complexation 122 Table 26. Formation Constants of DB30010-Cs+ Complex in Nonaqueous Solvents at Different Temperatures. SolvenSolvent Temp. (°C) Log Kf Nitromethane 70 3.65:0.01 60 3.70:0.02 15 3.99:0.10 3O 1.3020.05 10 1.67:0.11 0 > Acetonitrile 77 2.85:0.07 60 3.01:0.05 15 3.20:0.08 30 3.39i0.09 18 3.19:0.10 Acetone 60 3.1010.08 15 3.96:0.07 30 1.31:0.11 15 1.92:0.20 Methanol 60 3.36:0.08 15 3.70:0.05 30 1.18:0.07 15 1.65:0.02 0 >5 Pyridine 85 3.52:0.02 65 3.81:0.05 50 1.1320.01 30 1.1110.10 10 1.81:0.07 123 u— IO _ Ian v 9 .— P)’ NM 8" MeOH . Ac 7 #- Ahk l l 2.5 3.0 3.5 IO'3/ T Figure 19. Van't Hoff Plots for Complexation of Cs+ Ion by Dibenzo-30-Crown—10 in Various Solvents. 121 Table 27. Thermodynamic Parameters for the Complexation of Cs+ Ion by Dibenzo-30-Crown-10 in Various Solvents. AG°(30°C) AH° AS° Solvent (kcal/mole) (kcal/mole) (cal/mole °K) Nitromethane -5.97i0.07 -7.95:0.39 —6.66il.25 Acetonitrile -1.71i0.13 -5.1310.28 -1.5310.89 Acetone -5.5020.10 -l3.1810.51 r26.19il.69 Methanol -5.8liO.10 -l2.72iO.31 —22.82¢1.11 Methanola -5.77 -11.2 —18.2 Pyridine -6.1320.11 -7.9110.36 -5.93tl.13 aReference 32. 125 in various solvents increases in the order of acetone < methanol < nitromethane < pyridine < acetonitrile. In all cases the complexes are enthalpy stabilized but entropy destabilized. It should be noted that Similar behavior was previously observed in nonaqueous solutions. For example, in the case of the cryptate C222 exclusive complex with cesium 1 ion, the AH° and AS° values are -l2.9 kcal mole' and 1 and -13.7 e.u. 1 -26.8 e.u. in acetone, -8.6 kcal mole- in propylene carbonate and —5.7 kcal mole' and -11.2 e.u. in N,N-dimethylformamide solutions (96). Entropy destabiliza- tion was also observed by Izatt 33 31. for the complexa- tion of sodium and potassium ions by benzo-lS-crown—S and 18—crown-6 in water—methanol mixtures (31) and by Kauffmann gt_a1. for the complexation of potassium and rubidium ions by cryptand C221 and Na+, I<+, Rb+, and Cs+ ions by cryptand C222 in aqueous solutions (131). In the last case, it was assumed that the decrease in entropy was largely due to the rearrangement of the water structure upon the metastasis of a small inorganic cation into a large hydrophobic organic cation. While this explanation is quite feasible for aqueous solutions, it cannot be carried over to much less structured organic solvents used in this study. It seems reasonable to assume that the main reason for the negative entropy of the complexation is the decrease 126 in the conformational entropy of the ligand upon the formation of a metal complex. Large macrocycle ligand such as DB3OC10 should be rather flexible in the free state. The degree of flexibility would vary with the solvent, i;g;, with the extent of ligand-solvent inter- action. The formation of a rigid three-dimensional complex should decrease the conformational entrOpy of the ligand and thus, perhaps, give rise to negative entropy of the complexation. At the present time, however, thermodynamic data on the formation of macrocyclic complexes in non- aqueous solvents are quite sparse. Additional work is very necessary before the entropy destabilization of macro- cyclic ligands in nonaqueous solvents can be explained satisfactorily. 1.3. 002108 COMPLEXES WITH Cs+ In order to study the thermodynamic behavior of the complexation reaction between cesium ion and dibenzo-21— crown-8, extensive chemical shift-mole ratio studies were made of the Cs+-DB21C8 system in nitromethane, acetonitrile, acetone, methanol, and pyridine solutions over a wide temperature range. The measured chemical Shifts for dif- ferent ligand to metal ion mole ratios at various tempera- tures are given in Table 28 and the corresponding mole ratio plots are shown in Figures 20-21. In all systems Table 28. Cesium-133 Chemical Shifts of 0.005M 0s+ Ion 127 in the Presence of DB21C8 at VariouS'Tempera- tures. Solvent: Nitromethane DB21C8/05+ Temperature, °C 20 30 15 60 75 90 0.00 56.87 57.57 58.17 61.15 62.16 61.55 0.21 52.16 53.31 51.32 56.18 57.11 59.12 0.60 15.09 15.62' 16.61 17.61 18.88 51.06 0.70 13.51 11.38 11.91 16.17 17.59 18.82 0.91 10.13 11.37 12.29 13.16 15.09 16.87 1.05 39.60 10.59 11.55 12.70 11.21 11.55 1.10 39.31 10.12 10.90 12.38 11.00 15.00 1.18 39.11 39.86 10.59 11.83 13.19 11.13 1.10 38.12 39.01 39.66 10.66 11.79 13.38 1.87 38.03 38.58 39.29 39.97 10.78 11.98 2.39 37.96 38.12 39.07 39.71 10.28 11.60 2.88 37.96 38.12 39.00 39.53 10.11 11.15 Solvent: Acetonitrile 002108/Cs+ Temperature, °C 5 30 10 50 60 75 0.00 -37.35 -33.55 -32.08 -30.15 -29.29 -27.27 0.25 -26.65 23.21 -22.51 -21.11 -20.15 -19.52 0.62 -8.97 -7.50 -7.26 -7.12 -7.03 -6.57 0.72 -1.17 -3.70 -3.17 -3.85 -3.69 -3.62 0.81 0.72 -0.20 0.56 0.10 0.09 —0.21 0.98 6.16 5.85 5.00 1.36 3.71 2.73 1.09 9.25 8.32 7.51 6.77 6.16 5.69 1.13 9.88 8.55 7.93 7.08 6.96 5.90 1.38 12.51 12.12 11.73 10.61 10.80 9.87 1.78 13.30 13.68 13.05 12.93 12.81 11.57 2.32 13.11 11.11 13.90 13.85 13.75 13.67 2.78 13.11 11.22 13.95 11.10 11.21 11.35 128 Table 28. Continued. Solvent: Acetone DB21C8/03+ Temperature, °C 5 15 30 10 55 0.00 15.61 17.80 19.27 20.30 22.12 0.25 18.78 20.25 21.11 21.81 21.06 0.60 22.13 23.12 23.82 21.56 26.32 0.77 23.98 21.51 . 25.15 25.13 26.92 0.98 26.16 26.53 26.69 27.22 27.91 1.12 27.39 27.60 27.70 28.21 28.18 1.25 28.01 28.21 28.16 28.50 29.03 1.13 28.16 28.15 28.18 29.06 29.10 1.81 28.17 28.68 28.79 29.60 29.72 2.19 28.20 28.72 29.02 29.83 30.10 2.66 28.21 28.81 29.01 29.98 30.29 Solvent: Methanol DB21C8/03+ Temperature, °C 10 20 30 10 50 65 0.00 11.11 15.31 16.15 17.00 17.83 18.90 0.20 12.10 13.56 11.76 15.12 16.13 17.10 0.51 39.81 10.70 11.50 12.63 13.57 11.72 0.78 38.21 39.11 10.03 10.77 11.70 12.79 0.91 37.28 38.13 39.02 39.77 10.77 11.93 1.02 37.12 37.80 38.63 39.15 39.87 11.62 1.18 36.56 37.20 37.87 38.15 39.21 10.17 1.31 36.35 36.65 37.39 38.16 38.67 10.06 1.60 36.00 36.15 37.09 37.90 38.26 39.16 1.76 35.96 36.35 36.92 37.60 37.83 38.81 2.07 35.85 36.25 36.79 37.30 37.13 38.20 2.78 35.85 36.20 36.70 37.02 37.10 37.80 129 Table 28. Continued. Solvent: Pyridine DB21C8/08+ Temperature, °C 30 15 50 75 90 0.00 23.99 26.93 30.50 31.22 37.71 0.38 22.82 25.21 27.63 30.66 33.85 0.56 22.11 21.56 26.63 29.26 31.80 0.83 21.75 23.22 ‘ 21.69 27.17 29.81 0.92 21.36 22.83 21.25 26.70 28.72 1.05 21.19 22.13 23.68 25.61 27.87 1.26 20.98 21.59 22.77 21.07 26.21 1.52 20.88 21.28 22.60 23.16 21.92 2.03 20.81 21.20 22.36 23.01 21.25 2.39 20.79 21.18 22.28 22.88 21.13 2.83 20.79 21.13 22.20 22.80 23.96 130 as NITROMETHANE 45 - OC F :1 40" so 1 4s 20 35 - . , .1 I 14 11 L 1 0.0 0.5 In 1.5 2.0 2.5 3.1 [0824C81/[CJ] Figure 20. Cesium-133 Chemical Shifts vs. [DB21C8]/[Cs+] Mole Ratio in Nitromethane 56 Different Tem— peratures. 20 Figure 21. 131 - ACETONITRILE J J L 0.0 0.5 l 0 1.5 . zfo 2.5 [0824C8]/[C§ ] Cesium-133 Chemical Shifts E. [DB21C8]/[CS+] Mole Ratio in Acetonitrile at Different Tem- peratures. 10 132 30 - 33 / an a“"’ 15 ACETONE 25 8 PPM 20 Is ‘ I I I I I 0.0 0.5 III 1.5 2.0 2.5 [D824C81/[Cffl Figure 22. Cesium-133 Chemical Shifts vs. [DB21C8]/[Cs+] Mole Ratio in Acetone at Different Tempera- tures. 10 133 0 5 METHANOL 45 ‘ 5 ppm 40 - °c 55 A 30 :.2 A 13 35 - I L J . 1 0-0 0.5 1-0 1.5 2.0 2.5 3.0 [0824C81/[CJ] Figure 23. Cesium-133 Chemical Shifts vs. [Dszucej/[Cs+] Mole Ratio in Methanol at Di?ferent Tempera- tures. 134 35 30 Ppm 25 PYRIDINE O) K I 3 no 75 50 45 se— {o—————-030 20 I I J I J an &5 L0 L5 LB 15 10 [0324C81/ [CQ‘ ] Figure 2D. Cesium-133 tures. Chemical Shifts 15. [D32u08J/ECS+] Mole Ratio in Pyridine at Different Tempera- 135 studied, the curvature in the mole ratio plots is more pronounced at lower temperatures. This behavior implies that the complex formed is more stable at lower tempera- tures and that, of course, the complexation reaction is exothermic. The calculated formation constants of DBZDCB-Cs+ complex in different solvents at various temperatures are listed in Table 29. It is seen that such as with DB3OC10 the stability of the complex drastically increases as the temperature is lowered. Plots of fin Kf XE' the reciprocal of absolute temperature gave straight lines in all solvents used (Figure 25) and the AH° and AS° values were determined in the usual manner for the slope and the intercept of the plots. The calculated thermodynamic parameters for the complexation in various solvents are listed in Table 30. The data clearly show that in all cases studied the complexa- tion reaction is enthalpy stabilized but entropy destabi- lized. While the stability of DB2llD8-Cs+ complex (or AG°) is only marginally dependent on the nature of the solvent, the changes in enthalpy and particularly in the entropy of the complexation are very solvent dependent. The entropy change for the complexation in various solvents decreases in the order of pyridine > nitromethane > acetonitrile > methanol > acetone. Izatt and coworkers have studied the thermodynamics of the complexation of the cesium ion with DB2HC8 in 70% 136 Table 29. Formation Constants of DB2H08-CS+ Complex in Nonaqueous Solvents at Different Temperatures. Solvent Temp. (°C) Log Kf Nitromethane 90 3.37:0.05 75 3.52:0.03 6O 3.68:0.0h MS 3.91:0.03 30 h.11i0.08 20 U.26i0.06 Acetonitrile 75 3.19:0-03 6O 3.u5eo.ou 50 3.5710.05 no 3.77:0.07 30 3.9hio.07 5 M.50i0.05 Acetone 55 3.07:0.03 no 3.37:0.07 30 3.71:0.09 15 H.15i0.10 5 u.37:o.o7 Methanol 65 2.86:0.06 50 3.11:0.0u HO 3.36:0.10 3O 3.65:0.05 20 3.85:0.06 10 h.0hi0.10 Pyridine 9O 3.27:0.09 75 3.uuio.o7 60 3.60:0.0u us 3.76:0.08 30 u.OOi0.03 137 IIb 10‘ 9 .— {2‘ .E 8 .- NM PY AN 7— MeOH J l l l L as 23 33 12 3A 16 103/T Figure 25. Van't Hoff Plots for Complexation of Cs+ Ion by Dibenzo-2u-Crown-8 in Various Solvents. 138 Table 30. Thermodynamic Parameters for the Complexation of Cs+ Ion by Dibenzo-2u-Crown-8 in Various Solvents. AG°(30°C) AH° AS° Solvent (kcal/mole) (kcal/mole) (cal/mole °K) Nitromethane -S.7liO.ll -6.25i0.10 -l.7910.3u Acetonitrile -5.h710.10 -8.12i0.16 -8.6610.U8 Acetone -5.15i0.13 -ll.20t0.52 -20.08il.7l Methanol -5.06:o.07 -9.87io.46 —l6.lo:l.u9 Pyridine -5.55i0.05 -5.97i0.16 -l.39i0.u6 139 methanol solution by calorimetric titration. They also found negative values for the changes in the enthalpy and the entrOpy of the reaction (i.e., AH° = -8.09 kcal l and AS° = -lu.l e.u.). Since dibenzo-Zh-crown—lo mole- is a rather flexible ligand in free state, it is reason- able to expect that a change in conformation of the ligand, from a "loose" structure in the free state to a "rigid" structure in the complex, contributes to the large negative entropy changes of the reaction as a dominant factor. u.u. DB2lC7 COMPLEXES WITH Cs+ The thermodynamics of the complexation of the cesium ion with dibenzo-21-crown—7 was investigated in nitro- methane, acetonitrile, acetone, methanol, and pyridine by cesium-133 NMR. The variation of the chemical shift with the changing temperature was studied for the complexation reaction. At each temperature the cesium-133 chemical shift was monitored as a function of DBZlC7/Cs+ mole ratios. The data are given in Table 31 and the mole ratio plots are shown in Figures 26-30. As the temperature increases, the mole ratio plots show less curvature indicating the for- mation of a weaker complex at higher temperatures. This trend is evidence of the existence of an exothermic reac- tion between the cesium ion and the ligand. The calculated formation constants show the same trend (Table 32). 1“0 Table 31. Cesium-133 Chemical Shifts of 0.005M’Cs+ Ion in the Presence of DB21C7 at Various Tempera- tures. Solvent: Nitromethane DB2lC7/Cs+ Temperature, °c 15 30 “5 60 75 90 0.00 56.95 57.89 60.86 63.38 6“.39 65.81 0.26 “9.65 50.“3 53.01 5“.75 56.08 57.57 0.50 “2.0“ “3.37 ““.79 “6.12 “8.16 “9.33 0.77 33.57 35.22 36.86 38.51 “0.00 “2.59 0.9“ 29.33 30.75' 32.5“ 3“.59 36.23 38.58 1.11 25.80 27.37 29.“9 31.13 33.72 35.68 1.28 23.68 25.“l 27.05 28.78 30.82 33.02 1.“8 23.“5 2“.86 26.27 28.07 30.0“ 31.92 1.80 23.1“ 2“.2“ 25.72 27.06 28.“7 30.27 2.25 23.06 2“.15 25.02 26.67 27.1“ 28.55 2.67 23.05 2“.07 2“.95 26.27 26.90 28.00 Solvent: Acetonitrile + DB21C7/Cs Temperature, °C 20 30 “0 50 60 75 0.00 -33.2“ -31.93 -30.05 -28.““ -27.“3 -2“.9l 0.35 -25.96 -2“.79 -23.02 -21.86 -20.91 -19.13 0.6“ -l9.99 -l9.06 -l7.7“ -l6.72 -l6.02 -l“.79 0.82 -l6.33 -15.“9 -l“.63 -l3.70 -l3.3“ ~12.“7 0.92 -l“.32 -l3.“7 -l2.6“ -l2.23 -ll.53 -ll.l5 1.0“ -l2.30 -11.76 -ll.00 -10.76 -10.1“ —9.98 1.22 -9.98 -9.“7 -8.73 -8.35 —7.83 -7.57 1.33 -9.7“ -9.36 -8.27 -7.80 -7.“2 -7.19 1.50 -9.67 -9.05 -8.05 -7.“9 -6.96 -6.“9 1.83 -9.“5 -8.7“ -7.73 -7.11 —6.25 -5.71 2.52 -9.20 -8.52 -7.“2 -6.“1 -5.6“ -5.00 3.21 -9.1“ -8.“0 -7.18 -6.15 -5.27 -“.30 1“l Table 31. Continued. Solvent: Acetone DB21C7/Cs+ Temperature, °C 10 20 30 “0 55 0.00 16.67 17.86 19.“8 21.26 22.7“ 0.30 13.83 1“.98 16.“6 18.16 19.56 0.56 11.10 l2.“3 13.82 15.60 17.16 0.79 9.02 10.26 11.65 13.35 15.22 0.89 8.01 9.“0' 10.72 12.58 l“.06 1.03 7.00 8.32 9.67 11.3“ 13.36 1.12 6.37 7.5“ 8.9“ 10.80 12.58 1.20 6.15 7.39 8.78 10.65 12.35 1.“3 6.06 7.23 8.“2 10.13 11.73 1.90 6.00 7.07 8.2“ 9.67 ll.“2 2.37 5.98 7.03 8.20 9.“5 11.27 3.16 5.96 6.99 8.10 9.35 10.90 Solvent: Methanol DB2lc7/Cs+ Temperature, °C 20 30 “0 50 60 0.00 “5.“3 “6.97 “8.76 “9.92 50.50 0.30 38.13 39.53 “1.01 “1.85 “2.55 0.5“ 31.85 32.86 3“.3“ 35.65 36.82 0.68 28.60 29.92 31.07 32.55 33.86 0.90 23.17 2“.6“ 26.66 27.75 29.60 1.03 21.00 22.“0 2“.65 25.81 27.82 1.1“ 20.30 21.55 23.“0 2“.88 26.82 1.28 l9.“5 20.62 22.16 23.32 25.26 1.“6 19.00 19.90 21.39 22.“7 2“.33 1.85 18.83 19.66 20.82 21.70 23.10 2.“2 18.60 19.53 20.“5 21.23 22.80 3.0“ 18.53 19.3“ 20.30 21.00 22.“0 1“2 Table 31. Continued. Solvent: Pyridine DB2lCZ/Cs+ Temperature, °C 30 “0 50 60 0.00 2“.53 27.08 29.18 31.90 3“.92 38.50 0.17 19.80 22.““ 2“.77 27.18 30.20 31.97 0.50 11.12 12.83 15.07 17.“0 20.19 23.06 0.77 3.91 5.61, 7.2“ 9.10 12.20 l“.00 0.91 -0.67 1.03 2.“3 “.1“ 7.39 10.93 0.99 -2.30 -0.60 0.“9 2.89 5.5“ 8.“9 1.1“ -5.“7 -3.77 -2.68 -0.75 1.55 6.00 1.29 -5.63 -3.93 -3.39 -l.60 0.95 3.““ 1.51 -5.78 -“.63 -3.93 -2.30 -0.36 1.66 1.88 -5.82 -5.06 —“.50 -2.69 -1.07 1.0“ 2.37 -5.9“ -5.16 -“.72 -3.07 —1.36 0.15 2.97 -6.05 -5.32 -“.96 -3.“0 -2.03 -0.61 l“3 NITROMETHANE so so - 6 Ppm 4o - an - .C so 75 so 45 so 15 0.0 0.5 {a is in 275 3.0 [banal/[Cf] Figure 26. Cesium-133 Chemical Shifts 13. [DB2107]/[Cs+] Mole Ratio in Nitromethane at Different Tem- peratures. 1““ 351 -‘-‘75 4360 ~—£50 1_._,l *rOZO ACETONITRILE -35 h- I l i 1 L l 80 05 L0 L5 to 25 30 [0321014qu Figure 27. Cesium-133 Chemical Shifts vs. [DB21C7]/[Cs+] Mole Ratio in Acetonitrile at Different Tem- peratures. 1“5 25‘ ACETONE 20 I5 - 5 ppm °c ~—055 H)- ——o40 ——030 420 as: at e—clo 5.. l l J I l l 00 05 l 0 l 5 20 25 3 0 [banal/[CU Figure 28. Cesium-133 Chemical Shifts vs. [DB2107]/[Cs+] Mole Ratio in Acetone at Different Tempera- tures. 1“6 501 ‘ METHANOL 40- 5 ppm so - °c r-OBO A 50 20" ' *r'40 + —o 30 ‘—'—F* ——020 l J l 1 J I 50 05 10 L5 21| Z5 10 [0521C71/[C9‘] Figure 29. Cesium-133 Chemical Shifts vs. [DB21C7]/[Cs+] Mole Ratio in Methanol at DIfferent Tempera- tures. 1“7 PYRIDINE 30 20 '- 8 PP"‘ 10 '- _ °C 0 so r-o75 50 50 1 1 I 4 3L 0.0 0.5 1.0 1.5 2.0 2.51 3.0 [ 0521C 7] /[Cs*] Figure 30. Cesium-133 Chemical Shifts vs. [DB2107]/[Cs+] Mole Ratio in Pyridine at Different Tempera— tures. l“8 Table 32. Formation Constants of DBZlC7-Cs+ Complex in Nonaqueous Solvents at Various Temperatures. Solvent Temp. (°C) Log Kf Nitromethane 90 3.21:0.03 75 3.39:0.05 60 3.66:0.0“ “5 3.81:0.05 30 “.1“io.07 15 “.“0i0.10 Acetonitrile 75 3.15:0.05 60 3.“1i0.03 “0 3.73:0.05 30 3.95:0.0“ 20 “.11i0.05 Acetone 55 3.36:0.06 “0 3.6“:0.03 30 3.93:0.06 20 “.l9i0.0“ 10 “.52:0.0“ Methanol ' 60 3.5“:0.0“ 50 3.68:0.02 “0 3.83:0.02 30 3.96:0.06 20 “.1“i0.02 Pyridine 90 3.39:0.05 75 3.5710.07 60 3.78:0.05 50 3.89:0.0“ “0 “.07:0.05 30 “.27i0.07 l“9 Plots of 1n Kf vs, 1/T for the five systems are shown in Figure 31. The enthalpies and the entropies of the complexation were obtained in the usual manner from the slopes and the intercepts of the plots and the results are listed in Table 33. As is seen, in all cases the com- plexes are enthalpy stabilized but entropy destabilized. The enthalpy and entropy values vary very significantly with the solvent, but in all cases they compensate each other resulting in a nearly identical free energy for the complexation. The sequence of the increase in the entropy change in different solvents is acetone < acetonitrile < nitromethane < pyridine < methanol. “.5. DISCUSSION A deeper understanding of the thermodynamics of the complexation of the metal ions with macrocyclic crown ethers can be provided by the evaluation of the enthalpy and the entropy changes of the reaction. The magnitudes of enthalpy values are indicative of the metal ion-ligand interaction providing information about the type and the 'number of binding sites. The magnitude of the entropy values are indicative of the solvent-solute interactions. It can supply information about the relative degrees of solvation of the particles involved (lLSLfi the metal ion, the ligand, and the complex), the less of degrees of freedom of the ligand upon complexation, and the Charge 150 IO— 9.— O ¥Hu 15 0 8- MeO P)’ Ac NM AN 7 I l I l 2.8 3.0 3 3.2 3.4 3.6 IO/T Figure 31. Van't Hoff Plots for Complexation of Cs+ Ion by Dibenzo—Zl—Crown-T in Various Solvents. 151 Table 33. Thermodynamic Parameters for the Complexation of Cs+ Ion by DBZlC7 in Nonaqueous Solvents at Different Temperatures. Solvent (figzlggggg) (kc533;51e> (cal/figie °K) Nitromethane -5.75i0.10 -7.61i0.25 -6.25i0.78 Acetonitrile -5.“9i0.05 -8.2“i0.20 -9.1910.66 Acetone -5.“6i0.08 -11.13i0.“1 —18.6“il.35 Methanol -5.50i0.08 -6.6liO.1“ -3.5“iO.““ Pyridine -5.93i0.10 -7.22i0.18 -“.38i0.56 152 types involved in the reaction. In all three complexation reactions studied in this thesis, 1,24, complexation of the cesium ion with DB30010, DB2“C8, and DB2107, the contributions of the metal ion solvation and its charge type to the entropy changes in the same solvants are the same. Whereas, because of the difference in the size and in the number of the binding sites of the ligands, the contributions of the conformational change of the ligand and of the solvation of the ligand and the complex to the entropy changes are different. Thus in order to interpret the thermodynamic data, these factors must be considered. In all cases studied, the complexes are enthalpy stabilized but entropy destabilized. The entropies of the complexation of the cesium ion with DB30010, DB2“08, and DB21C7 in various solvents are compared in Table 3“. The solvation of the free and of the complexed cesium ion could be very different so that the complexation can in- fluence the structure of the solvent which could contribute to the entropy of the complexation. Since the solvents we worked with are less structured than water, the contribu- tion of this factor to the entropy would not be very im- portant. However, at the present time we have no knowledge about the extent of solvation of either the ligands or the complexes. A more important contribution to the negative entropy 153 Table 3“. Entropies of the Complexation of Cesium Ion by DB3OC10, DB2“C8, and DB21C7 in Various Solvents. AS°(ca1/mole °K) Solvent DB3OC10-Cs+ DB2“C8-Cs+ DB21C7-Cs+ Nitromethane -6.66il.25 -1.79i0.3“ -6.25i0.78 Acetonitrile -1.53i0.89 -8.66i0.“8 -9.19i0.66 Acetone -26.19i1.69 -20.08il.71 -l8.6“i1.35 Methanol -22.82¢1.ll -l6.lOil.“9 -3.5“i0.““ Pyridine -7.9“10.36 -5.93il.13 -“.38:0.56 15“ values would be the change in conformation of the ligands, from a flexible structure in the free state to a more rigid form, upon complexation. It is interesting to note that, with the exception of acetonitrile, in all solvents used, first, the entropy values of the DB2“C8°Cs+ complex are about “ to 6 e.u. less negative than those for DB3OC10-Cs+ complex, and second, the sequence of the solvent effect on the entropy values is the same for both complexes. According to these results, it seems reasonable to assume that among the various factors contributing to the negative entropies of the complexation of the cesium ion with DB30010 and DB2“C8, the decrease in the conformational entropy of the ligand upon complexation is the dominant one. Large macrocyclic ligands such as DB3OC10 and DB2“C8 are rather flexible in free state. The degree of flexibility would vary with the size of the ligands and with the solvent (Elia: with the extent of ligand-solvent interaction). DB30010 is more flexible than DB2“C8 in free state because of its larger size, and also DB3OClO-Cs+ complex is probably more rigid than DB2“C8-Cs+ complex because of its complete "wrap around" structure (Chapter 3). Therefore, the more negative entropy values for DB3OC10-Cs+ complex than those for DB2“C8-Cs+ are not unexpected. Acetonitrile as solvent, however, is an exceptional case probably because it can form a complex with the ligands, such as that reported with 18-crown-6 (135), and, therefore, can deviate the .-___.-___—--— " 155 results from the expected trend. 0n the other hand, in the case of the complexation of the cesium ion with DB2lC7, neither the sequence of the solvent effect on the entropy values nor the trend of the entropy values, i;gL, decreasing the entropy with increasing the size of the ligand, agree with the DB3OC10 and DB2“C8 cases. Since the cesium ion, with diameter of 3.68 A (128), has a very close size to that of DB2lC7 cavity, with the size of 3.“-“.3 K (23), it seems reasonable to assume that the DB21C7-Cs+ complex has a two-dimensional structure. Therefore the large conformation changes involving the ligand "wrapping around" the cation such as one observed of DB30010 are not expected in this ligand. Thus, unlike the cases of DB30010 and DB2“08, the conformational change of the ligand upon complexation is not necessarily the dominant factor in this case. Such an example can be found in the literature. Dibenzo—30-crown-10 (32) and 18-crown-6 (62) form complexes with the potassium ion in methanol solution with about the same entropy values of -l7.7 e.u. and -17.3 e.u. respectively. If the decreased conforma- tional change of the ligand upon complexation was the dominant contribution to the entropy values in both cases, we would expect the dibenzo-30-crown-10°Cs+ complex to have a much more negative entropy value than 18-crown-6-Cs+. It has been shown that potassium ion forms a three-dimen- sional "wrap around" complex with dibenzo-30-crown—10 (26,33) but a two-dimensional one with l8-crown-6 (136). CHAPTER 5 LITHIUM-7, SODIUM-23, CESIUM-133, AND THALLIUM-205 NMR STUDY OF Li+, Na+, 03+, and Tl+ ION COMPLEXES WITH 1,10-DIAZA-18-CROWN-6 IN VARIOUS NONAQUEOUS SOLVENTS 156 5.1. INTRODUCTION The macrocycle 1,10-diaza-18-crown-6 was first synthesized by Dietrich gt_a1. (137). 1.10-DIAZA-Is-CROWN-s Frensdorff (“) has shown that the substitution of two nitrogens for two oxygens in 18-crown-6 reduces the affin— ity of the ligand for alkali metal ions while enhancing the stability of the complexes of transition metal ions of the same size. Thermodynamic parameters for the complexation of some alkali earth and transition metal ions with this ligand in aqueous solution have been reported by Andereg (138) who used pH-metric titration for the measurements. Structural properties of some metal ion complexes with 1,10-diaza-l8-crown-6 in solution have been studied by proton NMR (l39,1“0) and the crystalline structure of the isolated complexes of the ligand with copper (II) and potassium ions have been determined (l“l,l“2). Copper (II) was shown to be located inside the cavity of the macro- cycle and it is bonded to the two nitrogen atoms and to the two oxygen atoms. The potassium is bonded to the four oxygen atoms of the ring in the same plane and to the two 157 158 nitrogen atoms from the top and the bottom of the macro- cycle plane. The present work was undertaken to determine the stoichiometry and the formation constants of 1,10-diaza- 18-crown-6 complexes of Li+, Na+, 05+, and Tl+ ions in various nonaqueous solvents. 5.2. RESULTS Lithium-7, sodium-23, cesium-133, and thallium-205 chemical shifts were measured as a function of the ligand/metal ion mole ratio in various solvents and the results are given in Tables 35-38. In all cases only one population average resonance of the metal ion was observed. Generally, this is only possible if a fast exchange exists between the two sites ($LEL, free and complexed ion) whose rate is larger than J27wAv (Av is the difference between the resonance frequency of each site). The chemical shift- mole ratio plots for different metal ions are shown in Figures 32-35- 5.2.1. 1,10-Diaza-18—Crown-6 Complexes with Li+ The frequency of the lithium-7 resonance in DMF, DMSO, and TMG was found to be independent of the ligand/lithium ion mole ratio (Figure 32). This behavior shows that the immediate environment of Lithium ion is not changed upon 159 Table 35. Mole Ratio Study of 1,10-Diaza-18-Crown-6 Com- plex with O-O2M.L1010“ in Various Solvents at 30°C. NM AN PC Ac + + + + L/Li 5(ppm) L/Li 5(ppm) L/Li 5(ppm) L/Li 5(ppm) 0.00 0.u3 0.00 2.56 0.00 0.63 0.00 —1.00 0.31 0.28 0.31 1.8“ 0.35 0.u3 0.39 -0.87 0.69 0.06 0.76 0.81 0.69 0.26 0.77 -0.73 0.88 -o.07 0.88 - 0.56 0.90 0.12 0.90 -0.69 1.00 -0.10 1.02 0.27 1.0“ 0.09 0.99 -0.67 1.10 -0.12 1.11 0.20 1.23 0.08 1.1“ -0.6& 1.33 -0.10 1.33 0.16 1.38 0.05 1.23 -0.62 1.60 -0.11 1.77 0.18 1.66 0.05 1.73 -0.52 2.1“ -0.11 2.u1 0.16 1.88 0.0“ 2.18 -0.u8 3.36 -0.11 3.2“ 0.16 2.31 0.06 3.“3 =0.u5 2.85 0.05 3.““ 0.05 DMF DMSO TMG Py + + + + L/Li 5(ppm) L/Li 5(ppm) L/Li 5(ppm) L/Li 5(ppm) 0.00 -0.63 0.00 1.03 0.00 -0.u5 0.00 -2.“1 0.38 -0.62 0.38 1.0“ 0.38 -o.uu e.uu -2.2u 0.75 -0.6u 0.75 1.0“ 0.75 —0.u6 0.67 -2.16 0.86 -0.62 0.87 1.06 0.88 -0.“7 0.91 -2.06 1.02 —0.63 1.08 1.0“ 1.00 -0.u5 1.02 -2.01 1.11 -0.62 1.16 1.06 1.11 -o.uu 1.10 -1.98 1.22 -0.63 1.u0 1.03 1.23 -0.“6 1.32 -l.93 1.57 -0.6“ 1.66 1.05 l.“3 -0.“6 l.“3 -l.88 2.07 -0.63 1.99 1.05 1.90 -0.uu 2.05 -1.66 3.13 -0.6“ 3.35 1.0“ 2.90 -0.u5 3.10 -1.u3 Id 160 Table 36. Mole Ratio Study of 1,10-Diaza—18-Crown—6 Complex with 0.05M.NaBPhu in Various Solvents at 30°C. NM Ac + + L/Na 6(ppm) Avl/2(Hz) L/Na 6(ppm) Avl/2(Hz) 0.00 1“.2l 13 0.00 6.93 18 0.32 11.97 11“ 0.“1 7.52 “1 0.70 9.“6 201 0.80 8.1“ 67 0.91 7.89 260 0.90 8.26 68 1.02 7.31 273 1.00 8.29 73 1.19 6.69 268 1.11 8.32 76 l.“7 6.88 27“ 1.22 8.32 79 1.8“ 6.83 276 1.“6 8.35 83 2.“8 6.79 276 2.11 8.“0 85 2.97 6.75 298 3.10 8.56 93 DMSO Py L/Na+ 5 Av (Hz) L/Na+ 5 Av (Hz) (ppm) 1/2 (ppm) 1/2 0.00 0.0“ “6 0.00 -1.15 21 0.“O 0.5“ 55 0.29 1.“3 55 0.75 0.86 68 0.70 “.81 10“ 0.88 0.90 67 0.86 6.70 121 0.96 1.02 7“ 0.9“ 7.08 137 1.09 1.11 7“ 1.09 7.““ 135 1.23 1.20 78 1.23 7.62 137 1.“9 1.26 80 1.55 7.70 1“6 2.02 1.50 96 2.09 7.66 1““ 2.68 1.81 105 2.57 7.67 155 161 ma.:mu um.» mm.mmn mm.s sm.mmu om.s mo.mmn os.s os.Hm- ew.a am.Hmu om.m aa.mmu mo.m :m.mmu mm.» Hm.san mm.m os.omu sm.m mm.omu nm.m mo.mmn mm.: mm.oan mm.m so.man mm.H H~.man m=.m mm.wan om.m H=.:u mm.H mm.sau ma.H s:.mau mm.m ms.smu ma.~ mm.m m=.H we.mau om.H mo.eau mm.a we.mmu =m.a mm.a mm.a as.:=- mo.H NH.HHu mm.a eo.smu sa.a Hm.m ea.fl ms.men no.0 mH.mu o=.H mm.mo- mm.H OH.OH mm.o ms.men mm.o mo.m om.a sm.ms- mo.H Hm.mH mm.o mm.m=u ms.o me.a mo.a am.msu mm.o =m.mH ms.o mo.mmu mm.o mm.m ma.o no.2mn No.0 os.ma mm.o ea.mmu m=.o mm.mH mm.o sa.mm- am.o am.mm mm.o mH.omu om.o sm.oa mm.o mo.ooau oo.o om.am oo.o He.mmu 00.0 mm.:m oo.o Asaavm +mo\q Aeddvo +m0\q AEqavm +mo\q Aeaavo +m0\q ooze mom ozg .ooom om mucosaon aaoaso> ca soH no spas onoEoo mucsonoumauousaouOH.H eo aospm ofioom oaoz .sm oases 162 .300000.030.0o .20000_ma0.00 00.03: 00.0 00.00- 00.0 03.00: 03.0 00.03- 00.3 30.00- 00.3 03.33- 00.0 00.00: 00.3 03.03- 03.0 30.00: 00.0 00.00: 03.0 00.0- 03.3 00.30: 00.0 00.03: 00.0 03.00: 00.3 00.00: 00.0 00.0- 03.0 00.03- 30.0 00.03: 00.0 00.00- 30.0 00.00- 03.0 00.0: 00.0 00.03- 33.0 00.00: 00.3 03.00- 30.0 00.00- 00.0 00.3: 03.0 00.30: 00.3 00.00: 30.3 00.00: 00.3 00.00- 00.0 00.3: 00.3 00.03- 00.3 30.00: 00.3 00.00-: 00.3 30.00: 00.3 00.0- 00.3 00.0: 00.3 00.30: 30.3 03.003: 03.3 30.00: 03.3 00.0: 00.3 00.0- 03.3 00.03: 00.0 30.003- 00.0 30.00: 30.3 00.0: 00.0 00.0- 00.0 30.03: 00.0 00.303: 00.0 00.00- 00.0 00.3- 30.0 00.0: 00.0 00.0: 00.0 «0.003-. 30.0 00.00- 00.0 00.3: 00.0 33.3 00.0 03.0 33.0 00.303: 00.0 03.00- 03.0 00.0- 00.0 03.0 03.0 00.00 00.0 03.003- 00.0 00.00: 00.0 03.0: 00.0 00.03 00.0 Asaavo +m0\q Aeaavm +mo\q Aegavm +mo\q “agave +m0\q Aegavo +mo\a 000 0020 00020 0020 .000030000 .00 03000 163 Table 38. Mole Ratio Study of 1,10-Diaza-18-Crown-6 Com- plexes with T1+ Ion in Various Solvents at 30°C. NMa ANb Aca + + + L/Tl 6(ppm) L/Tl 6(ppm) L/Tl 6(ppm) 0.00 36“.9 0.00 213.6 0.00 171.3 0.93 273.6 0.85 “5.1 0.50 112.8 1.12 251.8 1.05 12.0 0.70 83.“ 1.36 252.9 1.19 12.0 0.88 66.8 1.52 251.3 1.60 12.2 1.12 32.“ 1.86 251.9 1.91 12.8 1.32 31.9 2.67 251.8 2.81 12.“ 1.62 32.0 “.26 251.2 “.03 12.8 1.85 32.1 6.05 251.9 7.17 12.“ 2.82 31.9 “.15 32.“ 7.“l 32.6 DMFb DMSOb Pyb + + + L/Tl 5(ppm) L/Tl 5(ppm) L/Tl 5(ppm) 0.00 -129.9 0.00 -32“.1 0.00 -u.9 0.36 -85.1 0.“1 -230.8 0.36 -l3.l 0.8“ -10.9 0.72 -l72.2 0.83 -25.u 1.00 1“.“ 0.88 -l“7.5 0.91 —27.0 1.21 “0.1 1.00 -l30.9 1.01 -29.“ 1.“8 60.1 1.09 -11“.7 1.10 -30.6 1.7“ 66.0 1.20 -108.7 1.16 -30.8 2.80 71.3 1.55 -77.0 1.56 -3l.7 5.u8 72.7 1.85 -53.2 1.89 -31.8 2.9“ -l9.0 2.90 -3l.6 5.29 13.1 “.67 -31.3 7.80 21.“ 6.58 -30.9 a0.01M'T1C10u. b0.02110ch0“. l6“ 2 - DMSO I L *3 8 AN (PPm) a PC 0 - —0NM WMGr c 0.0 0:5 7:0 {.5 2:0 25 5.0 3.5 [0A18C51/[LV] Figure 32. Lithium-7 Chemical Shifts yg. [DAl8C6]/[Li+] Mole Ratio in Different Solvents. 165 l4 5 2' DMSO l L *1 0.0 6.5 LO 1:5 20 2.5 3.0 [0A15C51/[No*] Figure 33. Sodium-23 Chemical Shifts vs. [DAl8C6J/[Na+] Mole Ratio in Different SoIvents. 166 DMF fin TMO 0 in" aAN gyso 109 0.0 to 2.0 3.0 40 5.0 6.0 [DAIBC61/[Cf] Figure 3“ . Mole Ratio in Different Solvents. 7.0 8.0 Cesium-133 Chemical Shifts E. [DAlBC6]/[Cs+] 167 l J 0.0 LO 2.0 3.0 4.0 ~ 50 ab 7.0 8.0 [DAIBCOJIUI’] Figure 35. Thallium-205 Chemical Shifts 15. [ransom/[n+3 Mole Ratio in Different Solvents. 168 addition of the ligand, evidence of formation of a very weak complex at the best. These are solvents with high donicities and strong solvating abilities and, therefore can compete with the ligand for the lithium ion. On the other hand, in solvents of weak and medium donor ability such as nitromethane, acetonitrile, propylene carbonate, and acetone (with respective Gutmann donor numbers of 2.7, 13.1, 15.1, and 17.0) the lithium-7 chemi- cal shift is strongly affected by addition of the ligand, indicating existence of an interaction between the lithium ion and the ligand. As is seen in Figure 32, the lithium-7 resonance has a linear paramagnetic shift in nitromethane, acetonitrile, and propylene carbonate (diamagnetic in acetone) upon addition of the ligand which begins to level off at mole ratio of about 1, indicating formation of a complex with 1:1 stoichiometry between the lithium ion and the ligand. A gradual diamagnetic shift was observed in pyridine solutions which could be attributed to the formation of a weak complex. The limiting chemical shifts and the formation constants of the complexes were obtained by computer fitting the chemical shift-mole ratio data to an equation (discussed in Chapter 3) which relates the observed chemical shift to the complex formation constant using KINFIT program. The results are given in Table 39. While the cavity size of the macrocycle is larger than the size of the lithium ion, the resulting complexes seem 169 Table 39. Formation Constants and the Limiting Chemical Shifts of l,lO-Diaza-l8-Crown-6-Li+ Complexes in Various Solvents. Solvent Log Kf 611m(ppm) Nitromethane >5 -O.lli0.02 Acetonitrile 4.39:0.Nl 0.16:0.00 Propylenecarbonate 3.67:0.25 0.05:0.00 Acetone - 2.13:0.08 -0.36t0.02 Dimethylformamide ~0.0 -—-- Dimethylsulfoxide No.0 ---- Tetramethylguanidine ~0.o ---— Pyridine O.h3:0.08 0.6t0.3 170 unexpectedly strong, particularly in the solvents with low donicity. The formation constants are even greater than those reported for lB-crown-6 - Li+ complex in the same solvents (81). The stability of the complexes in different solvents decreases in the order of nitromethane > acetonitrile > propylene carbonate > acetone > pyridine which is the order of increase in Gutmann donor number of the solvents. Despite the close range of the limiting chemical shifts in nitromethane, acetonitrile, propylene carbonate, and acetone, chemical shift of the complexed lithium ion is solvent dependent, indicating incomplete insulation of the cation from the solvent by complexation. 5.2.2. l,lO-Diaza—18-Crown-6 Complexes with Na+ Sodium-23 NMR study of the sodium ion complexes with the ligand in a number of solvents such as propylene car- bonate and DMF was limited by the quadropolar broadening of the 23Na resonance (Avl/2)500 Hz) which makes the precise measurements of the chemical shift impossible. The change in 23Na chemical shift upon addition of the ligand to the sodium salt solution in acetonitrile was found to be less than 0.5 Ppm which is in the range of the error of the chemical shift measurement and, therefore, cannot be used in such studies. Thus, the complexation of sodium ion with the ligand was studied only in nitromethane, acetone, DMSO, and pyridine solutions, where the linewidths 171 are narrow enough to measure the chemical shifts accurately. The addition of the ligand to the sodium tetraphenyl- borate solution in the above four solvents results in an upfield or a downfield chemical shift which begins to level off at mole ratio of about 1 indicating formation of a 1:1 complex between the cation and the ligand. The formation constants and the limiting chemical shifts for the complexes are shown in Table H0. The limiting chemical shifts of the complex in nitromethane, acetone, and pyridine are close to each other which possibly indicates that the cation is mostly covered by the ligand. In DMSO solutions, however, the complex seems to be more solvent dependent. It should be noted that DMSO has a strong solvating ability and, therefore, can compete with the ligand for the cation. With the exception of pyridine, the complex formation constant increases with decreasing Gutmann donor number of the solvents. The results in pyridine are even un- expected because it should be a good solvating solvent as indicated by the magnitude of its sodium-23 chemical shift (1H3) and the high Gutmann donor number of 33.1. This is possibly because of existence of nitrogen atom as a soft donor which cannot strongly solvate a hard ion such as sodium. The sodium ion complexes with l,lO-diaza-l8- crown-6 are much weaker than those with lB-crown-6 in the same solvents (1&4) because of the introduction of two nitrogen atoms into the lB-crown-6 ring. According to the 172 Table “0. Formation Constants and the Limiting Chemical Shifts of l,lO-Diaza-18-Crown-6-Na Complexes in Various Solvents. Solvent Log Kf 611m(ppm) Nitromethane 3.37:0.13 6.7Hi0.02 Acetone 1.96:0.18 8.83:0.09 Dimethylsulfoxide 1.19:0.08 2.86:0.21 Pyridine ' H.12i0.30 7.69:0.02 173 Pearsons Hard-Soft-Acid—Base (HSAB) theory (1&5), the‘ interaction of the sodium ion as a hard acid with the nitrogen atom as a soft base should be weaker than that with the oxygen atom as a hard base. Another interesting point to note is that the line width of the sodium-23 resonance at half height (Avl/z) increases almost linearly with increasing the ligand to sodium ion mole ratio, indicating creation of a more un— symmetric environment of 23Na nucleus because of the complexation, and begins to level off after mole ratio of 1, showing a 1:1 stoichiometry for the complex (Table 36). 5.2.3. 1,10-Diaza-18-Crown—6 Complexes with Cs+ In all solvents used, with the exception of DMSO, the addition of the ligand to the cesium ion solution produces a large but gradual paramagnetic or diamagnetic shift of the cesium-133 resonance (Figure 34). In none of the solvents used does the chemical shift of the com- plexed cesium ion reach a limiting value even at ligand to metal ion mole ratio of about eight. The results seem to indicate formation of a weak 1:1 complex between the cesium ion and the ligand. Using cesium-133 NMR, Mei 22.2l- (9H) have shown that cesium ion forms both 1:1 and 2:1 (ligand to metal) com- plexes with l8-crown-6. In pyridine, acetone, DMF, and 17“ PC solutions they observed a downfield shift of 13303 resonance followed by a relatively sharp break at the mole ratio of l and an upfield shift which gradually ap- proaches a limiting value. This behavior was explained by formation of a 1:1 complex followed by addition of a second molecule of the ligand to form a 2:1 sandwich com- plex. It is reasonable to assume that the large cesium ion will also form 1:1 and 2:1 complexes with l,lO-diaza- 18-crown-6 which has a cavity of about the same size as l8-crown-6. However, no clear evidence of the formation of a 2:1 complex between cesium ion and l,lO—diaza-18- crown-6 was observed in this study. The formation constant and the limiting chemical shifts for 1:1 complexes of the cesium ion with the ligand in different solvents are given in Table “1. A large difference in the limiting chemical shifts of the complexes (about 65 ppm) in different solvents clearly indicates that the cesium ion is too large to fit into the ligand's cavity, therefore the cation mostly remains under the influence of the solvent. The complexes are much weaker than the 1:1 cesium ion complex of 18-crown-6 reported by Mei gt al, (9“). This is simply because of the exist- ence of the two nitrogen atoms as a soft base in the ligand which cannot strongly interact with the cesium ion, known as a hard acid (1&5). Therefore, the resulting complex should be weaker than that with 18-crown-6. This instability 175 Table H1. Formation Constants and the Limiting Chemical Shifts of l,lO-Diaza-lB-Crown-S-Cs+ Complexes in Various Solvents. Solvent Log Kf 611m(ppm) Nitromethane 2.79:0.02 -28.06iO.ll Acetonitrile 2.26:0.02a -55.93¢o.08 Acetonitrile 2.30:0.01b -56.50so.02 Propylene Carbonate 1.95:0.02 -36.7620.36 Acetone 1.89:0.01a —h3.88i0.l9 Acetone l.92:0.01° -u5.56:0.l7 Dimethylformamide 0.61:0.07 -6l.O 18.6 Dimethylsulfoxide m0.0 _ ---- Trimethyleneoxide l.9fli0.02 1.91:0.72 Tetramethylguanidine 1.55:0.02 -79.8210.38 Pyridine 2.62:0.01 -52.D2:0.06 amoniesscm. b0.03I_vI_tsSCN. CO.OlM_CsBPhu. 176 indicates that if the 2:1 complexes are formed, the second formation constant would be «K1 and, therefore, probably would not show up on the mole ratio plots. It is seen that the complex stability is not affected either by a change in the concentration of the salt, in acetonitrile solutions, or by changing the anion, in acetone solutions. It is evident therefore, that at low concentrations of the cesium salts, used in this study, the formation of the complex is unaffected by ion pairing. It is also interesting to note that, with the exception of pyridine solution, the stability of l,lO-diaza-lB-crown- 6 ° Cs+ complex increases by decreasing the donicity of the solvent. 5.2.“. l,lO-Diaza-18-Crown-6 Complexes with T1+ In nitromethane, acetonitrile, acetone, and pyridine solutions the thallium-205 resonance shifts nearly linearly with the ligand to metal ion mole ratio until a mole ratio of 1 is reached. Further addition of the ligand does not have any further effect on the 205T1 resonance (Figure 35). This behavior indicates formation of a strong complex (Kf > 105) between Tl+ ion and the ligand in the above solvents. On the other hand, in DMSO and DMF, solvents with high donicity, a gradual diamagnetic shift of the thallium—205 resonance was observed upon addition of the ligand which tends to level off after the mole ratio of 177 1, indicating formation of weaker complexes in these Sol- vents. The formation constants and the limiting chemical shifts for the complex in various solvents are shown in Table 42. The formation constants of the complexes in nitromethane, acetonitrile, acetone, and pyridine are greater than 105 and their precise values could not be determined by our technique. In DMSO and DMF, solvents with high donor ability, the stability of the complex is still significant. The scattered values of the complex limiting chemical shifts shows that the solvent molecules can still interact with the complexed thallium ion, pos- sibly from the top and the bottom of the ligand's plane. 5.3. DISCUSSION The formation constants of Li+, Na+, 03+, and Tl+ ion complexes of 1,10-diaza-lB-crown-6 and 18—crown-6 in various solvents are compared in Table H3. The most probable ionic diameters, calculated by using a combination of deduction from rO values and experimental electron density maps by Ladd (128), for the above cations are 1.72 A, 2.2a A, 3.68 A, and 3.08 A, respectively. The cavity size of the ligand is about 2.6-3.2 A (23). The stabilities of the l,lO-diaza—lB-crown-6 complexes decreases in the following 4. order Tl+ > Li+ > Na > Cs+. Among the various factors contributing to the complex stability, the two following 178 Table U2. Formation Constants and the Limiting Chemical Shifts of l,lO-Diaza-l8-Crown—6-Tl+ Complexes in Various Solvents. Solvent Log Kf 611m(ppm) Nitromethane >5 251.8 Acetonitrile >5 12.4 Acetone >5 32.4 Dimethylformamide - 3.22:0.08 7U.0i0.3 Dimethylsulfoxide 2.1310.03 “0.6i1.l Pyridine >5 -31.5 179 . QG Ouflmhfluflflv . QQH ”Ur—OH 0H0-U .HQ OUGOHOUUM—D . NQH QUGOHUHOKO nx ex so.os~o.o ma Ho.oa~o.~ on.oa~H.« mo.cans.o nx H.nn an nun- nun- o.oa .nu- ~o.oann.H a.-- o.oe u..- u--- use un.n u--- n-.. gnu: ~o.oa«o.a u-.. u-.. n.-- n--- oza «o.ow«c.n so.owa«.a o.c;. Ho.ow~a.a 9.08. no.6waH.H c.c€ nc.cHnH.~ «.mm Gaza ma.cwa.n «o.o«m~.~ lulu wo.owmn.n no.cwao.o null c.ce. we.ow-.n c.o~ man mA «A No.0won.fl nA Ho.onom.a oa.oucm.u mo.cwna.u nA o.sa . o< ma.ow«u.« «A Ha.owmo.~ null No.owna.a III: n~.ow~o.n lulu H.ma on «A o~.owom.m «c.ow«n.~ mA ~o.owc~.u nail H«.owmn.« mA H.«H z< III! «A «A mA No.c«ms.u mu.cwun.n an mA h.~ :z v+uo o+nz A+wn o+HH +10 +uz +AA +H9 2: uoo>Hom oassosouaa osssosoumanosoaonofi.a «a mo; .mu=o>aom maowuo> :« claaououwa new olozou0ImHIonoualoH.H mo moonano com +HH new +mo .+wz .+Aq mo mucmuoaoo nouuoauom .n« manna 180 factors are especially important: first, the relative sizes of the cation and the ligand, and second, the strength of the interaction between the cation and the coordination sites of the ligand. In order to interprete the data, these factors should be considered. In the case of the cesium ion both factors are against the complexation. Cesium ion is a hard acid (1&5) which cannot strongly interact with the soft nitrogen atoms of the ligand. It is also too large to fit inside the ligand's cavity. Therefore, the cesium ion complex in different solvents is the weakest one in the series. Despite the hard character of the sodium ion and consequently its weak interaction with the soft nitrogen atoms of the macro— cycle ring, this cation has a suitable size for the ligand. Thus it is not surprising that l,lO-diaza-18-crown-6 forms stronger complex with sodium ion than that with cesium ion. Although the size of the lithium ion is a little smaller than that of the sodium ion for the ligand and it is harder than the sodium ion, it forms an unexpectedly more stable complex with the ligand than sodium ion. Among the alkali metal ions, lithium ion is known to have a great tendency toward covalent bond formation because of its great polarization power (1A6). Therefore, the existence of such covalent bond in the lithium ion complex with the ligand could possibly result in such a great stability of the complex. In the case of the thallium (I) 181 ion both factors are in the favor of the complexation and, therefore, the thallium (I) ion complex is the strongest among the series. The substitution of two nitrogen atoms for two oxygen atoms in 18-crown-6 greatly influences the complexation of the ligands with the metal ions, as shown by the data in Table “3. The sodium and the cesium ion complexes are weakened appreciably by the nitrogen substitution. This is Just as expected: as the negative charge on the co— ordination sites drops by the substitution, the electro- static interaction between the ligand and the cation would diminish and the resulting complex would be weaker. The effect is more pronounced in the case of the cesium complex (iLEL, weakening of the complex by about two orders of magnitude) than that of the sodium complex. This is simply because the cesium ion is a harder acid than the sodium ion which results in a weaker interaction between the substituted nitrogen atoms, as a soft base, and the cesium ion than that with the sodium ion. The effect of the nitrogen substitution on the lithium ion complex is exactly the opposite: the stability of the complex is greatly increased. Because of its small size and large polarizability, lithium is known to have some chemical behavior that resembles the chemistry of magnesium (146). Because of this unusual behavior, lithium ion has a great ability to form covalent bond. It is 182 evident, therefore, that the increased stability of the lithium ion complex upon the substitution is because of the existence of both electrostatic and covalent bonding, in contrast with the cesium and the sodium ion cases where the bonding is only electrostatic in nature. In the case of the thallium (I) ion complex, in most of the solvents used the stability of the complexes are not exactly cal- culated and, therefore, they cannot be compared. we can expect, however, the thallium (I) ion complex with l,lO- diaza-l8-crown-6 to be more stable than that with 18-crown- 6, such as that observed in DMSO solution. This is be— cause the interaction of the soft thallium (I) ion with the soft nitrogen atoms of the ring is quite strong. CHAPTER 6 A STUDY OF DYNAMICS OF CESIUM ION COMPLEXES WITH DIBENZO-30-CROWN-10, DIBENZO-ZH-CROWN-8, AND DIBENZO-Zl-CROWN—7 IN ACETONE AND METHANOL 183 6.1. INTRODUCTION While there are several scientific reports available in the literature on the study of thermodynamics of the complexation of alkali ions by crown ethers, the complexa- tion kinetics of the cyclic polyethers have received less attention. Kinetic studies of alkali metal complexation with crown ethers are usually impeded by several factors: complexes are not usually strong enough and therefore must be studied at high metal ion concentrations; the reaction rates are usually high and experimental difficulties are' encountered especially if work at high concentrations is required; and finally, the complexes are usually color— less, so that spectrophotometric measurements of rates are rarely possible. In an early sodium-23 NMR study of complexation of sodium ion by dibenzo-l8-crown-6 in DMF solution, Shchori _t _1. (87) concluded that the strong ionic strength dependence of the kinetics indicates that the dynamic equilibrium involves interaction of solvated sodium ion with uncoordinated ligand rather than a simple bimolecular exchange equilibrium. The same authors (88) studied the effect of solvents and aromatic ring substituents on the kinetics of complexa- tion of sodium ion by dibenzo-l8-crown-6. They used the sodium-23 NMR technique and worked with DMF, methanol, 18“ 185 and dimethoxyethane solvents. They found that the activa- tion energy for decomplexation of sodium ion by dibenzo- 18-crown-6 is the same in all solvents (”12.6 kcal/mole) but is considerably less (8.3 kcal/mole) for dicyclohexyl- l8-crown—6 in methanol. Shporer and Luz (93) have used 39K and 87Rb NMR to determine the rate constants for potas- sium and rubidium ion decomplexation by dibenzo-l8-crown-6 in methanol solution at low temperatures. They found that the decomplexation of rubidium is much faster than of potassium ion. Using spectrophotometric measurements, Chock (32) determined the rate constants for complexation of several monovalent cations by dibenzo-30-crown-10 in methanol solution. From the resulting data he suggested the exist- ence of a fast crown ether conformational transition preceding the complexation reaction: CR 1 + K12 + fast M + M K: MCR CR2 21 in which CR1 and CR2 denote different conformations of the uncomplexed ligand. Eyring and coworkers (1A8,lu9) have determined the rate constants for cation decomplexation by l5—crown-5 and 18-crown-6 in aqueous solutions by ultrasonic 186 absorption method. They concluded that the slow decomplexa- tion of 18-crown—6-Cs+ complex is the reason for the high selectivity of l8-crown-6 for potassium ion over the other alkali ions. The purpose of the work described in this chapter is to study the kinetics of the complexation of the cesium ion by DB3OClO, DB2AC8, and DB21C7 in acetone and methanol solutions by cesium-133 NMR lineshape analysis at various temperatures. 6.2. DETERMINATION AND INTERPRETATION OF THE LINESHAPES The modified Block equations proposed by McConnell (150) which describe the motion of the X and Y component" of magnetization in the rotating frame are shown as fol— lows: dG A _ -1 —1 dt + GAGA ‘ '1YH1M0A + T13 GB ‘ TA GA (1) dG B _ . -1 -1 dt + O‘BGB ‘ ’ lYHlMoB + TA GA " TB GB (2) where dA = l/T2A - 1(wA-w) and GB = l/T2B - 1(wB-w) G = u + iv (3) 187 u and v are the transverse components (i;g;, absorption and dispersion mode lineshapes) of magnetization along and perpendicular to the rotating field, H1. The solutions of Equations (1) and (2) appropriate for slow passage are obtained by dG dG atfl=at§=0 (u) The equations can be solved for GA and GB. Noting that MOA = PAMO and MOB = PBM (5) O the total complex moment is T + T + T T (a P + a P ) = -1YH1MO A B A B A A B B_ (6) (1 + aAtA) (1 + aBTB) - l G=GA+GB Fast Exchange. In the limit of rapid exchange, TA and TB are small and Equation (6) reduces to G = -in M = - (7) The imaginary part is 2 (8) v=..'YH 0 2 2 v 1 + T2 (PAwA + P 1M w w) B B - 188 representing a resonance line centered on a mean frequency of wmean = PAwA + Paws (9) with a linewidth given by T17=TPL+TEE (10) 2_ 2A 2B If the exchange is not quite rapid enough to give complete collapse, the central signal centered on ”mean will appear to have a larger width than that given by Eq. (10). A corrected form of Equation (10) can be obtained by putting w = w in Equation (6) and expanding in mean powers of T. This gives an effective transverse relaxa- tion time P PB ._ A 2 2 2 _1_ T2 2A 2B where PA and PB are relative populations at sites A and B, respectively, the quantities “A and “B are the resonance frequencies at the two sites at a given temperature in the absence of exchange, and T2A and T2B are the respective relaxation times at each site at a given temperature. The lifetime of interaction is defined as 189 T = --——-- (12) If at a given temperature the I value is greater than /2 , where Aw = IwA - mBI, two separate resonance lines nAw are observed for the two respective sites. If the T value is less than :Eé’ only one population average resonance is observed. 1H NMR, where the trans- In some cases such as 13C and verse relaxation times are long enough, the following assumption has been made (151-153): fi§;.= Tl—.= 0 (13) 2A 2B Since cesium-133 is a nuclei with narrow natural line- width of about 1 Hz (quadrupole moment of q= 0.003 barns) and consequently long enough transverse relaxation time, it seems reasonable to assume that the above assumption is also valid for 13303 NMR case. In all cases studied, the ligand to cesium ion mole ratio was 0.5 and also the formation constants of the complexes were greater than 103 (Chapter 3). Therefore PA = PB = 1/2 (in) The relationship between the frequency in radians per second, w, and that in cycles per second, v, is defined 190 as w = 2flv (15) In general the relaxation time is given by (15“): ‘3; = rate of removal of molecule from 1th state by exchange Ti number of molecules in the ith state (16) If we consider the following complexation reaction 13f M + L + ML (17) kb then d /dt ?1_ = ._U—ML (18) B ML 0 /dt Ti -- MC (19) A M Since d ML _ ‘ ‘d‘t" kb CML (20) dM " Ht" = kf CMCL (21) 191 Therefore we have with Equation (23) For equal population that we used; Equation (1A) .=_1_ 2kb (22) (23) (2“) (25) (26) By substitution of Equations (12), (1A), (15), (2A), and (26) in Equation (11) we will come out with the following equation: It is known that (27) 192 g; = NW (28) where W is the linewidth at half height. By substitution of Equation (28) in Equation (27) we have 2 b - 2W (28) where VA and VB are the resonance frequencies of the two sites in the absence of exchange, and W = A033: - Aviig. The AvEBZ and Avis; are the linewidths at half height in the presence and absence of exchange, respectively. Thus above the coalescence temperature, the linewidth of the sample solution (with the ligand/cesium ion mole ratio of 0.5) was measured and corrected for natural broadening by subtracting the linewidth of the resonance signal of a solution of pure cesium salt in the same solvent at the same temperature from it. The resonance frequencies of the solutions containing free and complexed cesium ion, VA and VB, respectively, in the same solvent at the same temperature were also measured. The kb value at each temperature was then calculated from Equation (29). Slow Exchange. If the lifetimes TA and TB are suf- ficiently large compared with the inverse of the separa- tion (wA - mB)_l, the spectrum will consist of two distinct signals in the vicinity of the wA and wB frequencies. For example, if the frequency of w is close to “A, and thus far 193 away from ”8’ GB is effectively zero and the solutions of Equations (1) and (2) become P T A G s GA z inlMO ———5———— (31) 1 + “ATA The imaginary part is P T' A v = -in1MO ' 2 2A 2 (32) A broadened signal centered at “A with width given by parameter -1 l -1 Y = T2A T2A + TA (33) There will be a corresponding signal centered on “B“ This shows that the exchange leads to an additional broad- ening of the individual signals. If T2A is known, measure- ments of the width of these broadened signals provide a means of estimating TA' Thus below the coalescence temperature the kb value can be calculated from kb = WW (3“) where W is the corrected linewidth of the resonance signals, i.e., w = Avggg - Avggg. 19“ The Arrhenius activation energy, E is given by a, 81m k b 2 ( ) = E /RT (35) BT V or P a or . Ea The activation energy was obtained from the slopes of the activation plots, (i.e., 2n kb vs. 1/T) by using a linear least squares program. The thermodynamic parameters of activation were calculated from the following equations: # _ AHO — Ea — RT (37) # kT AH: ASO = R in kb - R in Tr + T (38) ¢_¢ a AGO — AHO - TASO (39) where AGé, AH:, and A8: are the standard free energy of activation, the standard enthalpy of activation, and the standard entropy of activation, respectively, and k and h are Boltzmann and Planck constants, respectively. The rate constants for the forward reactions (i.e., complexation) 195 were calculated from the following equation: kf = K kb (110) where K is the complex stability constant at 25°C. 6.3. RESULTS AND DISCUSSION In order to study the kinetics of the complexation re- actions between cesium ion and dibenzo-2l-crown-7, dibenzo- 2A-crown-8, and dibenzo-30—crown-10 in acetone and methanol solutions, the cesium-133 NMR spectra of the cesium thio- cyanate in the presence of the ligands (with a ligand/Cs+ mole ratio of 0.5) at various temperatures were obtained. The results are shown in Figures 36-Al. The required information in the absence of exchange (i.e. ”A: v8, and ref 1/2 NMR spectra of the solutions of salt (site A) and of the Au ) was obtained by lineshape analyses of the cesium-133 completely complexed cesium ion (site B), collected at exactly the same conditions as the exchange case. The rate constants for the release of the cesium ion from the complexes, kb, at temperatures above coalescence were calculated from Equation (29), below the coalescence temperature from Equation (3A). The data are listed in Table AA. The Arrhenius plots, log k vs. l/T are shown b in Figure A2. Activation energies (Ea), rate constants (kb), and values of AH:, ASS, and AG: for the release of -42 4000 2500 Hz .Figure 36. Cesium-133 NMR Spectra of 0.02 M CsSCN, 0.01 M DBZlC? Solution in Acetone at Various Tempera- tures. 197 Cs+ Cs C+ 0C -84 J 12 ‘_- -27 Figure 37. Cesium-133 NMR Spectra of 0.01: M CsSCN, 0.02 13 DB21C7 Solution in Methanol at Various Tem- peratures. Cs+ CsC -54 MMWM MVW -40 -20 4000 2500 Hz Figure 38. Cesium-133 NMR Spectra of 0.02 M CsSCN, 0.01'M DB2hC8 Solution in Acetone at Various Tem- peratures. 199 w-” o ._- -55 , -i_ 4000 2500 Hz Figure 39. Cesium-133 NMR Spectra. of 0.02 M CsSCN, 0.01 M DB2ACB Solution in Methanol at Various Tem- peratures. 200 J L -28 'I500 A zsunJHz1 Figure 40. Cesium-133 NMR Spectra of 0. 01 M CsSCN, 0. 005 M DB3OC10 Solution in Acetone at various Tem- peratures. 201 WM A‘WA’AA ANNA fiiflMALWMNAAWwWVA/A MA; .. ANAL, -au nflflfifimmquA AfihfiflhnMVUAJ%fiflqu~%vaWANNA LANA/AWAAVAWWA ”WW-MA AAA WMWMWWAT Figure Al. Cesium-133 NMR Spectra of 0. 01 M CsSCN, 0. 005 M DB30010 Solution in Methanol at Various Tem- peratures. 202 Table nu. Temperature Dependence of the Rate Constants for the Release of 03+ Ion from DB21C7-Cs+, DB2HC8 Cs+, and DB3OClO°Cs+ Complexes in Acetone and Methanol. Ligand Solvent Temp. (°C) log k b DB21C7 Acetone -82* 0.93 '76* 1.2“ -70* a -6u* a -57 2.18 -50 2.U8 -U2 2.77 Methanol -8H* 1.66 -72* 2.07 -65* 2.32 —61 2.35 -H8 2.85 -27 3.U3 DB2HCB Acetone -98* 0.75 -88* 1.36 —77* 1.8H -67* a -5u 2.84 -MO 3.30 -28 3.73 Methanol -92 1.93 —8U 2.24 -78 2.H7 -73 2.62 -55 3.32 203 Table uu. Continued. Ligand Solvent Temp. (°C) log k DBBOClO Acetone -80 1.84 -68 2.32 —56 2.66 -Ul 3.12 -28 3.28 Methanol -90* a —80* a -6M 2.95 -55 3.1M -u9 3.34 -h5 3.H2 -H2 3.50 -30 3.77 * Below the coalesence temperature. aThe k D values are not calculated. 201A l I l 4.0 4.5 5.0 5.5 6.0 103/T Figure 42. Arrhenius Plots of log kb vs. l/T for the Release of Cs+ Ion in Acetone and Methanol with Large Crown Ethers. A-DB3OClO'Cs+ in Methanol, B-DB2uC8-0s+ in Methanol, c—DB3001o-Cs+ in Acetone, D-DB2lc7-Cs+ in Methanol, E—DB2uc8-0s+ in Acetone, F-DB2lC7-Cs+ in Acetone. 205 the cesium ion from the complexes, as well as the rate constants for the complexation reactions (kf) are given in Table H5. It is immediately obvious that the nature of the solvent plays an important role in the kinetics of the complexation reactions. In all cases, the energy of activation for the release of the cesium ion from the complex in methanol is less than that in acetone. It should be noted that methanol is a solvent with higher donicity than acetone as expressed by their respective Gutmann donor number of 25.7 and 17.0 (106). The same effect was observed by Cahen gt_al. (78) for the release of lithium ion from cryptate C2ll.Li+ in pyridine, water, DMSO, DMF, and formamide solutions. By contrast Shchori et al, (88) found that the activation energy for decomplexation of sodium ion by dibenzo-18- crown-6 is the same in methanol, DMF, and dimethoxyethane. However, two of these solvents, i;g;, DMF and methanol, have about the same donicity, while the donicity of di— methoxyethane is not known. The activation energy for the release of the cesium ion from the complex is also dependent on the size of the ligand. In the same solvent, there is an inverse relation- ship between the activation energy for decomplexation of the cesium ion by the ligand and the size of the ligand (Table #5). These results seem to indicate that the major barrier to removal of the cesium ion by the complexes is the 206 .COHumH>op cpmocmpmp .ponss: Locop acmepzcm AH.oVN.HH Ao.avm.mmn A:.ovo.: Am.HVm.e Am.ovm.m Az.ovm.m oaoommo AH.ovs.oH Am.mvm.oan Am.ovs.m Ao.Hv:.z As.avs.m Am.ovm.m wozmmo AH.ove.HH Ae.Hve.eHu Am.ova.o Ao.0vm.m Am.ovm.m Am.ovs.m eofimmo s.mm Hocooooz AH.ovm.HH Am.va.omn Am.ovm.m Am.ovo.m A:.ova.m Am.ova.m oaoommm AH.ovo.OH AN.NVH.OHI Ao.ovo.e Aw.Hve.o As.va.oa Am.ovm.m moemmo AH.ovm.OH Am.mvm.sn Ae.ovs.w Am.HVH.s AH.HVm.m oAm.ovm.m aofimmo o.eH ooooooa mHoE\HmOx mHoE\HmOx oHoE\Hmox Hloom HIS Hlomm mHoE\Hmox pcmwfiq mzo uco>aom O O O .m Axommmv noa nmo xmo mIOonx agoaxox m .Hocmnpoz ocm oCOpoo< CH moonQEoo czoho mmpmq mEom Eopw COH no mo ommmfiom pom mpmuoemhmm OHEmsmUoELmne 6cm mmumm owcmnoxm .m: magma + 207 energy required to affect a conformational rearrangement. The lower activation energy for the release of the cesium ion from the larger crowns would then be a result of the greater flexibility of the ligand. In both solvents, the rate constants for decomplexation of the cesium ion by DB2AC8, kb, is greater than those for the release of the cation by DB3OClO and DB21C7. The rate constants for the decomplexation reactions seem to be determined by the rigidity of the resulted complexes. Cesium ion with diameter of 3.68 A (128) has a convenient size for the cavity of DB21C7 with the size of 3.u—u.3 A (23) to form a firm "cation in the hole" complex. We have already shown (Chapter 3) that DB3OC10 is large enough to form a rigid "wrap around" complex with the cesium ion, while due to differences in the sizes of DB2UC8 cavity and of the cesium ion either a complete "wrap around" or a "cation in the hole" complex. The rate constants for both complexation and decomplexation of the cesium ion by DB3OC10, i424, kf = 7.5 x lo8 M“Asec'l and kb = 3.5 x 10‘A sec-l, are in a satisfactory agreement with the values reported by Chock (32), iLEL, kf = 8 x 108 M.1 sec-1 and kb = u.7 x 10LA M‘1 sec‘l. As is seen in Table 45, while the enthalpy and the entropy of activation for the release of the cesium ion from the ligands are very sensitive to the solvent and to the size of the ligand, the free energies of activation 208 in all cases are about the same. Such a compensation of the enthalpy with the entropy is not an uncommon occurr- ence (155). It is interesting to note that the entropy of activation for decomplexation of the cesium ion by large crowns (Table #5) are in opposite direction with the corresponding overall entropies (Table 3”, Chapter M). This is in contrast with the entropy values reported for the release of the cesium ion from cryptate C222 in DMF solution by Mei §£.§l-'(95): where the activation and the overall entropies have the same sign and are close in value. From the results the authors concluded that the activated complex should resemble the final state, i424) the solvated cesium ion and the free cryptand. The op- g o and AS° values for decomplexation of posite sign for AS the cesium ion by large crowns, however, suggests that the conformational entrOpy changes may play an important role (156). Since the activation energy for decomplexation of the cesium ion by the ligands increases with increasing the donicity of the solvent, the transition state must involve a substantial ionic solvation. A sample entropy profile for decomplexation of the cesium ion by DB3OClO in acetone is shown in Figure 43. It clearly shows that the transi- tion state must be more ordered than both the initial and the final states, i.e., the solvated complex and the solvated cesium and the free ligand. Thus the possible 209 08300055“ A i AS°=26.2e.u. >~ 0. t g vDB30C10'Cf A5 =-46.5e.u. .5 A A$*=—20.3e.u. V , V (DBaoCm-Cf) ——> Reaction Coordinate Figure “3. Entropy Profile for the Release of Cs+ Ion From DB3OClO-Cs+ Complex in Acetone. 210 mechanism for the release of the cesium ion from the complex can be shown by # # AS AS DB3OClO°CS+ <—————- (DB3OC10’CS+)# 0 + 1 real root b2 a3 II 1f-+ 57 — 0 + 3 real roots b2 a3 III 7T' + 27 < 0 + 3 real roots Case I, x = A + B Case II and III, use trigonometric form 2“ — % Cos (% + 120°) 24 — % Cos (% + 2A0°) 221 Now, solve for [M] in (25). Then substitute in following equations, K C [M] [ML] = f L Kf[M] + 1 K C [M] [MA] = ip M Kip[M] + 1 sobs = XMGM + XMLGML + XMA‘SMA _ LE1. [ML] [ML] 6obs ' CM. 5M + CM 5ML + -Cfi—6MA Use final form of 6 in EQN subroutine. obs Coding symbols in EQN, a = AA p = PP b = BB q = QQ A = AAA r = RR B = BBB ¢ = FE y = R Cos ¢ = CFE CONST(l) = 1p CONST(2) = CM CONST(3) = 61p CONST(A) = 6M U(l) = SML XX(1) = 0L U(2) = Kf xx<2) = sobs (26) (27) (28) (29) nnnnnnnnn 2222 sUEonuTxNE En- C'Nucu «aunt.ITADEo.;TAo£.IwToLADoxthp.NopT.N VAR.~ouux.x.n.tTwa. Eons IBTA. TEST I.Av.9E§lD.IA°~ EPSolTYDoXXoRXTvpcnxl oFODan-Fu.°.ZIoTO E EDNA 210:025321g onTOL0“9IIJOYODY0VEC70NCSToCONSToNDAToJDAToHODToL0° T Eons VT corona/FDEDT f‘zuCN/DOIM D} "(NSION I UW'vx daox on: Zen“: *40K‘CCWJOD<'M "- ~Odlxn~3333490 3 WM —< q u—«o 'hn UKDU n [’2 Cflflx 444430" T 0 S : t F {nu PAID ¢ TA r upng UFTAL T F COMPLEX IGAND «IFT Ttla ) I -—,- vad W )0 —mv—-—wmmw30\ 0‘GO‘ LAoiis.ts.la IS weliéchAo 5.2:: 2i rggvATC 0 ALL THREE ROOTS FLUVKEDOT s SUDST TUTE cau-cn ATED “ETA couc. ~ on n DUAT n 30 0|9§5‘:Egggg?gfgl?:0‘285§;%;::é:;;gtléi%l{febtgb'wiitgou5TTItoo°cnus O QESYD’DELTA-XX(2T ' 0 2’ Rf ITJDN 3 cc~rluut 6 ZSerguE s CONTINUE éééyflttfloflE.-|! Go To 20 20 com [NU 8E1U PM E o ONTINUE Wu ‘3 ggfiflflufi ii cngINUE VIP" 12 ro' UxNu: EN REFERENCES 10. ll. 12. 13. 1“. 15. REFERENCES C. J. Pedersen, J. Amer. Chem. Soc., Q2, 7017 (1967). "Structure and Bonding", Vol. 16, Springer-Verlag, New York, NY,1963 A. I. Popov and J. M. Lehn, in "Chemistry of Macro- cyclic Compounds", G. A. Melson, Ed., Plenum Press, New York, New York, in press, Chapter 9. H. K. Frensdorff, J. Amer. Chem. Soc., 9;, 600 (1971). J. J. Christensen, J. 0. Hill, and R. M. Izatt, Science, 113, fl5a (1971). J.-M. Lehn and J. P. Sauvage, J. Amer. Chem. Soc., 21, 5700 (1975). N. K. Dalley, J. S. Smith, S. B. Larson, K. L. Mathe- son, J. J. Christensen, and R. M. Izatt, J. Chem. Soc. Chem. Commun., 8“ (1975). D. J. Cram et a1., Pure Appl. Chem., 3;, 327 (1975). K. H. Pannel, W. Yee, G. S. Lewandos, and D. C. Hambrick, J. Amer. Chem. Soc., 29, 1H5 (1977). R. Ungaro, E. E1 Haj, and J. Smid, J. Amer. Chem. Soc., gg, 5198 (1976). S. C. Shah, K. Kopolow, and J. Smid, J. Polym. Sci. 19., 2023 (1976). w. Steinmann and T. A. Kaden, Helv. Chim. Acta, 5Q, 1358 (1975). B. E. Jepson and R. Dewitt, J. Inorg, Nucl. Chem., 3_8_, 1175 (1976). N. Matsura g£_a1., Bull. Chem. Soc. JaEn., £9, 12H6 (1976). E. M. Arnett and T. C. Moriarity, J. Amer. Chem. Soc., 2;, 14908, (1971). 223 16. 17. 18. 19. 20. 21. 22. 23. 2H. 25. 26. 27. 28. 29. 30. 31. 32. 33. 224 T. H. Ryan, J. Koryta, A. Hofmanova—Matejkova, and M. Brezina, Anal. Letters, 1, 335 (1974). C. J. Pedersen and H. K. Frensdorff, Angew. Chem. Internal. Edit., 11, 16 (1972). J. J. Christensen, D. J. Eatough, and R. M. Izatt, Chem. Rev., 13, 351 (l97h). I. M. Kolthoff, Anal. Chem., 51, IR (1979). C. J. Pedersen, Fed. Proc., Fed. Amer. Soc. Exp. Biol.,{gl, 1305 (1968)? D. C. Tosteson, Fed. Proc,, Fed. Amer. Soc. Exp. Biol., 21, 1269 (1968). H. Lardy, Fed. Proc., Fed. Amer. Soc. Exp. Biol., g1, 1278 (1968). C. J. Pedersen, J. Amer. Chem. Soc., 92, 386 (1970). N. S. Poonia and M. R. Truter, J. Chem. Soc. Dalton, 2062 (1973). D. G. Parsons, M. R. Truter, and J. N. Wingfield, Inorg. Chim. Acta, 13, As (1975). M. A. Bush and M. R. Truter, J. Chem. Soc. Perkin 33. 3A5 (1972). M. Mercer and M. R. Truter, J. Chem. Soc. Dalton, 2u69 (1973). D. E. Fenton, M. Mercer, N. S. Poonia, and M. R. Truter, J. Chem. Soc. Chem. Commun., 66 (1972). J. D. Owen and M. R. Truter, J. Chem. Soc. Dalton, accepted for publication. M. Pinkerton, L. K. Steinrauf, and P. Dawkins, Bio- Chem. Biophys. Res. Commun., :5, 512 (1969). G. A. Rechnitz and E. Eyal, Anal. Chem., 33, 370 (1972). P. B. Chock, Proc. Nat. Acad. Sci. USA, 69, 1939 (1972). D. Live and S. I. Chan, J. Amer. Chem. Soc., 9Q, 3769 (1976). 225 3A. R. M. Izatt, R. E. Terry, D. P. Nelson, Y. Chan, D. J. Eatough, J. S. Bradshow, L. D. Hansen, and J. Jé Christensen, J. Amer. Chem. Soc., 98, 7626 (197 ). 35. A. Hofmanova, J. Koryta, M. Berzina, and M. L. Mittal, Inorg. Chim. Acta, gg, 73 (1978). 36. J.—M. Lehn, P. Vierling, and R. C. Hayward, J. Chem. Soc. Chem. Commun., 296 (1979). 37. M. Ciampolini, P. Paoletti, and L. Sacconi, J. Chem. Soc., 2994 (1961). 38. L. Sacconi, P. Paoletti, and M. Ciampolini, J. Chem. Soc., 5115 (1961). 39. P. Paoletti and A. Vacca, J. Chem. Soc., 5051 (196A). A0. L. Sacconi, Pu Paoletti, and M. Ciampolini, J. Chem. Soc., 50h6 (196“). A1. G. Schwarzenbach, Advan. Inorg. Chem. Radiochem., ;. 257 (1961). “2. G. Anderegg, Helv. Chim. Acta, 58, 1718 (1965). H3. G. Anderegg, Helv. Chim. Acta, 36, 1833 (1963). AN. G. Anderegg, Proc. 8th I. C. C. C. (Vienna) 3A (196“). 45. G. Anderegg, Helv. Chim. Acta, 31, 1801 (196“). “6. D. K. Cabbiness and D. w. Margerum, J. Amer. Chem. Soc., 91, 6540 (1969). U7. P. Paoletti, L. Fabbrizzi, and R. Barbucci, Inorg. Chem., 1g, 1961 (1973). A8. R. P. Hinz and D. W. Margerum, Inorg.vChem., 1;, 29u1 (197M). A9. A. Dei, and R. Gori, Inorg. Chim. Acta, 13, 157 (1975). 50. M. Kodama and E. Kimura, J. Chem. Soc. Chem. Commun., 325 (1975). 51. M. Kodoma and E. Kimura, J. Chem. Soc. Dalton, 116 (1976). 52. 53. 5“. 55. 56. 57. 58. 59. 60. 61. 62. 63. 6A. 650 66. 226 F. Arnaud-neu, M. J. Schwing-weill, J. Juillard, R. Louis, and R. Weiss, Inorg. Nucl. Chem. Letters, 13, 367 (1978). R. M. Izatt, J. H. Rytting, D. P. Nelson, B. L. Hay - more, and J. J. Christensen, Science 16h, AH3 (1969). R. M. Izatt, D. P. Nelson, J. H. Rytting, B. L. Haymore, and J. J. Christensen, J. Amer. Chem. Soc., 93, 1619 (1971). -—' P. U. Fruh, J. T. Clerc, and W. Simon, Helv. Chim. Acta, 99, INNS (1971). W. K. Lutz, P. U. Fruh, and W. Simon, Helv. Chim. Acta, 93, 2767 (1971). C. U. Zust, P. U. Fruh, and W. Simon, Helv. Chim. Acta, 2. A95 (1973). E. Schori and J. Jagur-Grodzinski, Isr. J. Chem., 11, 243 (1973). T. E. Hogen-Esch and J. Smid, J. Phys. Chem., 79, 233 (1975)- R. M. Izatt, R. E. Terry, B. L. Haymore, L. D. Hansen, N. K. Dalley, A. G. Avondet, and J. J. Christensen, J. Amer. Chem. Soc., 98, 7620 (1976). R. M. Izatt, J. D. Lamb, R. E. Asay, G. E. Maas, J. S. Bradshow, and J. J. Christensen, J. Amer. Chem. Soc., 99, 613A (1977). R. M. Izatt, J. D. Lamb, G. E. Maas, R. E. Asay, J. S. Bradshow, and J. J. Christensen, J. Amer. Chem. Soc., 99, 2365 (1977). E. Mei, J. L. Dye, and A. I. Popov, J. Amer. Chem. Soc., 99, 1619 (1976). E. Mei, J. L. Dye, and A. I. Popov, J. Amer. Chem. Soc., 99. 5308 (1977. R. M. Izatt, N. E. Izatt, B. E. Rossiter, and J. J. Christensen, Science, 199, 99A (1978). R. M. Izatt, J. D. Lamb, B. E. Rossiter, N. E. Izatt, and J. J. Christensen, J. Chem. Soc. Chem. Commun., 386 (1978). 67. 68. 690 70. 71. 72. 73. 7N. 75. 76. 77. 78. 79. 80. 81. 82. 83. SN. 227 R. M. Izatt, R. E. Terry, L. D. Hansen, A. G. Avondet, J. S. Bradshow, N. K. Dalley, T. E. Jensen, and J. J. Christensen, Inorg; Chim. Acta, 99, 1 (1978). F. Bloch, Phys. Rev., 19, N60 (l9N6). F. Bloch, W. W. Hansen, and M. Packard, Phys. Rev., £2. 37 (19u5). N. F. Ramsey, Phys. Rev., 11, 567 (1950). N. F. Ramsey, Phys. Rev., 19, 699 (1950). N. F. Ramsey, Phys. Rev., 99, 2N3 (1952). J. Kondo and J. Hamashita, J. Phys. Chem. Solids, 19, 2N5 (1959)- c. ggverell and R. E. Richards, Mol. Phys., 10, 551 (19 ). '— R. G. Barnes and W. V. Smith, Phys. Rev., 99, 95 B. Ditrich, J.-M. Lehn, and P. Sauvage, Tetrahed. Letts, 2885 (1969). E. T. Roach, P. R. Handy, and A. I. Popov, Inorg. Nucl. Chem. Letters, 9, 359 (1973). Y. M. Cahen, J. L. Dye, and A. I. Popov, J. Phys. Chem., 19, 1289 (1975). "“““ Y. M. Cahen, J. L. Dye, and A. I. Popov, J. Phys. Chem., 19, 1292 (1975). A. Hourdakis and A. I. Popov, J. Solution Chem., 6. 299 (1977). A. J. Smetana, Ph.D. Thesis, Michigan State University, East Lansing, MI (1979). D. H. Haynes, B. C. Pressman, and A. Kowalsky, Bio- chem., 19, 852 (1971). R. Bodner, M. S. Greenberg, and A. I. Popov, Spect. Letts. 9, N89 (1972)- A. M. Grotens, J. Smid, and E. deBoer, J. Chem. Soc. Chem. Commun., 759 (1971). 85. 86. 87. 88. 89. 90. 91. 92. 93. 9N. 95. 96. 97. 98. 99. 100. 101. 102. 228 R. H. Erlich, E. Roach, and A. I. Popov, J. Amer. Chem. Soc., 99, N989 (1970). J. Grandjean, P. Laszlo, F. Vogtle and H. Sieger, Angew. Chem. Internal. Ed., 11, 856 (1978). E. Shchori, J. Jagur-Grodzinski, Z. Luz, and M. Shporer, J. Amer. Chem. Soc., 99, 7133 (1971). E. Shchori, J. Jagur-Grodzinski, and M. Shporer, J. Amer. Chem. Soc., 99, 38N2 (1973). J. M. Ceraso and J. L. Dye, J. Amer. Chem. Soc., 22; N432 (1973). J. P. Kintzinger and J.-M. Lehn, J. Amer. Chem. Soc., 9_6_. 3313 (19711). J. M. Ceraso, P. B. Smith, J. S. Landers, and J. L. Dye, J. Phys. Chem., 91, 760 (1977). J. S. Shih, Ph.D. Thesis, Michigan State University, East Lansing, MI (1978). M. Shporer and Z. Luz, J. Amer. Chem. Soc., 91, 665 (1975). E. Mei, A. I. Popov, and J. L. Dye, J. Phys. Chem., §_1_, 1677 (1977). E. Mei, A. I. Popov, and J. L. Dye, J. Amer. Chem. Soc., 99, 6532 (1977). E. Mei, L. Liu, J. L. Dye, and A. I. Popov, J. Solu- tion Chem., 9, 771 (1977). F. J. Kayne and J. Reuben, J. Amer. Chem. Soc., 99, 220 (1970). C. Srivanavit, J. I. Zink, and J. J. Dechter, J. Amer. Chem. Soc., 99, 5876 (1977). R. W. Briggs and J. F. Hinton, J. Mag; Res., 99, 363 (1979). A. E. Martell, Adv. Chem. Ser., 99, 272 (1967). D. D. Traficante, J. A. Simms, and M. Mulcay, J. Mag. Res., 19, N8N (197N). D. A. Wright and M. T. Rogers, Rev. Sci. Instrum., 33. 1189 (1973). 103. lON. 105. 106. 107. 108. 109. 110. 111. 112. 113. 11N. 115. 116. 117. 1L18. 229 J. W. Cooper, "An Introduction to Fourier Transform NMR and the Nicolet 1080 Data System", Nicolet Instrument Corp., Madison, WI, 1972. D. H. Live and S. I. Chan, Anal. Chem., fig, 791 (1972). G. Foex, G. L. Gorter, and L. J. Smits, "Constantes Selectionees, Diamagnetisme et Paramagnetisme, Relaxation Paramagnetique", Massoy and Cie Editeurs, Patis, 1957. (a) V. Gutmann and E. Wychera, Inorg. Nucl. Chem. Letts, g, 257 (1955); (b) V. Gutmann, "Coordination Chemistry in Non- aqueous Solvents", Springer-Verlag, Vienna, 1968. M. S. Greenberg, Ph.D. Thesis, Michigan State Uni- versity, East Lansing, MI (197N). J. L. Dye and V. A. Nicely, J. Chem. Educ., 99, NN3 (1971). A. I. Popov, Pure Appl. Chem., 91, 101 (1979). Y. M. Cahen, Ph.D. Thesis, Michigan State University, East Lansing, MI (1975). W. J. Dewitte, Ph.D. Thesis, Michigan State University, East Lansing, MI (1975). E. Mei, Ph.D. Thesis, Michigan State University, East Lansing, MI (1977). E. G. Bloor and R. G. Kidd, Can. J. Chem., 99, 3N25 (1968). C. Deverell and R. E. Richards, Mol. Phys., 19, N21 (1969). A. L. Van Geet, J. Amer. Chem. Soc., 99, 5583 (1972). J. J. Dechter and J. I. Zink, J. Amer. Chem. Soc., 21. 2937 (1975). R. W. Briggs and J. F. Jinton, J. Solution Chem., .1. 1 (1978). A. I. Popov, Chapter 13 in "Solute-Solvent Interactions", J. F. Coetzee and C. D. Richhie Editors, M. Dekker Inc., New York, NY, Vol. , 1976, page 271. 119. 120. 121. 122. 123. 12N. 125. 126. 127. 128. 129. 130. 131. 132. 133. 13N. 135. 136. 137. 138. 230 J.-M. Lehn and J. Simon, Helv. Chim. Acta, 99, 1N1 (1977). J. F. Hinton and R. w. Briggs, J. Mag; Res., 22. 379 (1977). G. E. Maciel, 99.91., Inorg. Chem., 9, 55N (1966). R. H. Cox and H. W. Terry, J. Mag. Res., 19, 317 (197“)- R. H. Erlich and A. I. Popov, J. Amer. Chem. Soc., 23. 5620 (1971). M. S. Greenberg, R. L. Bodner, and A. I. Popov, 1. Phys. Chem., 11,_2NN9 (1973). R. Freeman, 99,91., Mol. Phys., 9, 75 (1959). W. J. DeWitte, 99_§1,, J. Solution Chem., 9, 337 (1977). A. G. Lee, Coord. Chem. Rev., 9, 289 (1972). M. F. C. Ladd, Theoret. Chim. Acta, 19, 333 (1968). J.-M. Lehn, Struct. Bonding (Berlin), 19, 1 (1973). N. Ahmad and M. C. Day, J. Amer. Chem. Soc., 99, 9N1 (1977). M. Shamsipur and A. I. Popov, J. Amer. Chem. Soc., 101, u051 (1979). JPM. Lehn, J. P. Sauvage, and B. Deitrich, J. Amer. Chem. Soc., 99, 2916 (1970). A. Hourdakis, Ph.D. Thesis, Michigan State University, East Lansing, MI (1978). E. Kauffmann, J.-M. Lehn,,and J. P. Sauvage, Helv. Chim. Acta, 99, 1099 (1976). D. J. Cram, J. Org. Chem., 99, 2NN5 (197N). M. R. Truter, Struct. Bonding, 19, 71 (1973). B. Dietrich, J.-M. Lehn, and J. P. Sauvage, Tetrahed. Letts., 99, 2885 (1969). G. Anderegg, Helv. Chim. Acta, 99, 1218 (1975). 231 139. B. Dietrich, J.-M. Lehn, and J. P. Sauvage, Tetra- hedron, 99, 16N7 (1973). 1N0. A. Knochel, J. Oehler, G. Rudolph, and V. Sinnwell, Tetrahedron, 99, 119 (1977). 1N1. M. Herceg and R. Weiss, Inorg. Nucl. Chem. Letts, Q; "35 (1970). 1N2. D. Moras, B. Metz, M. Herceg, and R. Weiss, Bull. Soc. Chim. France, 551 (1972). 1N3. M. S. Greenberg and A. I. Popov, J. Solution Chem., ,3, 599 (1975). INN. J. D. Lin and A. I. Popov, to be published. 1N5. (a) R. G. Pearson, J. Amer. Chem. Soc., 99, 3533 (1963). (b) R. G. Pearson, J. Chem. Educ., 99, 6N3 (1968). 1N6. F. A. Cotton and G. Wilkinson, "Advances in Inorganic Chemistry", Third Edition, Interscience Publisher, New York, NY, 1972, page 190. 1N7. G. Rounaghi and A. I. Popov, to be published. 1N8. G. W. Liesegang, M. M. Farrow, N. Purdie, and E. M. Eyring, J. Amer. Chem. Soc., 99, 6905 (1976). 1N9. G. W. Liesegang, M. M. Farrow, F. A. Vazquez, N. Purdie, and E. M. Eyring, J. Amer. Chem. Soc., 99, 32N0 (1977). 150. H. M. McConnell, J. Chem. Phys., 99, N30 (1958). 151. F. A. L. Anet, J. Amer. Chem. Soc., 99, N58 (196N). 152. J. E. Anderson and J.-M. Lehn, J. Amer. Chem. Soc., .82.. 81 (1967). 153. F. A. L. Anet and A. J. R. Bourn, J. Amer. Chem. Soc., .8_9. 760 (1967). 15N. Amdur and Hammes, "Chemical Kinetics Principles and Selected Topics", McGraw Hill 00., New York, NY, 1966, page 1N7. 155. K. U. Laider, "Chemical Kinetics", McGraw Hill 00., New York, NY, 1965, page 251. 232 156. J. L. Dye, "Progress in Macrocyclic Chemistry", Vol. 1, R. M. Izatt and J. J. Christensen, Editors, Wiley-Interscience, New York, NY, 1979, page 63.