w. Date 0—7639 July 18, LIBRARY Michigan State University This is to certify that the thesis entitled Physicochemical studies of magnesium complexes presented by Pierre-Henri C. HEUBEL has been accepted towards fulfillment of the requirements for Ph. D. degree in _§hemi§:rx flaw , Major professor Professor Alexander I. POPOV 1978 PHYSICOCHEMICAL STUDIES OF MAGNESIUM COMPLEXES By Pierre-Henri C. Heubel A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Chemistry 1978 ABSTRACT PHYSICOCHEMICAL STUDIES OF MAGNESIUM COMPLEXES By Pierre-Henri C. Heubel Magnesium-25 NMR measurements were made in a variety of solvents on several magnesium salts as a function of salt concentration. Magnesium chloride and bromide were prepared fOllowing the method described by Ashby and Arnott [redox reaction between magnesium turnings and a mercury (II) halide in tetrahydrofuran (THF)]. When ion pairing takes place in solution, a chemical shift change (up to 30 ppm in propylene carbonate [PC]) and a broadening of the NMR signal (up to 260 Hz in PC) are observed. The limiting chemical shift of magnesium-25 in different solvents was determined by studying solutions of MgClz, MgBrZ, MgClzI4 and MgBr214 in acetone, acetonitrile, dimethylformamide, dimethylsufoxide, methanol, PC, THF and water. Complexation of magnesium ions by the cryptand Cle in dry methanol results in a small line broadening, while the complexation by phosphonoacetate (PA) in water broadens the signal to such an extent that the NMR signal disappears. The larger distortion of the electric field by the ligand PA is thought to be responsible for this behavioral difference. Pierre-Henri C. Heubel Infra-red studies of acetonitrile solutions of MgBrz showed complexation of the magnesium ion by the solvent. The acidity functions of phosphonoacetic acid (PAA) and its complexation with magnesium ion were also investigated by other physicochemical techniques. The behavior of the phosphorus-31 and carbon-l3 NMR signals 1; pH, as well as the accompanying change in Jc_p coupling constant show evidence for a relationship between the acetic acid moiety of PAA and its second acid dissociation constant (0K2). The pKa's of PAA were studied potentiometrically by coulometric titration at ionic strengths from 0.l to 0.02 and numerical values were obtained by using the computer program MINIQUAD 76A. Tetra- methylammonium bromide was used as a supporting electrolyte. Thermodynamic pKa's of 2.0, S.ll i 0.04 and 8.69 i 0.05 were obtained by linear least squares fitting and extrapolation to zero ionic strength. The thermodynamic quantities for the second and third protonation of PAA were obtained by studying the temperature dependence of the formation constants. These two protonations of PAA are entropy and enthalpy stabilized (entropy stabilization strongly dominant) as shown below: 003 = -6.9 i 0.3 kcal/mole A03 = -ll.7 i 0.4 kcal/mole Aug = -0.2 i 0.3 kcal/mole AH3 = -l.3 i 0.4 kcal/mole A53 = 22.6 i 0.9 cal/mole, °K 053 = 35 i l cal/mole, °K In the case of K], the experimental results were not sufficiently precise for thermodynamic calculations. Pierre-Henri C. Heubel The stoichiometry of the completely deprotonated PAA-cadmium complexes was investigated by cyclic voltammetry. A l:l and a 2:l (ligandzmetal) complex were found. The respective formation constants of 8 x 103 and 8 x 101 were determined at 0.4 ionic strength. Different complexes (namely mono- and completely deprotonated ones) were observed by potentiometry in the case of alkali and alkaline earth metal ions. Thermodynamic values of the magnesium- PAA complexes were determined in the same manner as the thermodynamic pKa's. Log Kf values of 3.0 i 0.3 and 5.58 i 0.09 were obtained for the mono- and deprotonated complexes, respectively. The thermodynamic quantities: AGO = -7.2 i 0.7 kcal/mole AHO 3.0 i 0.7 kcal/mole 0 AS 35 i 2 cal/mole, °K were calculated from the temperature dependence of the deprotonated complex formation constant. The change in charge-to-size ratio of the different ions upon complexation is thought to be responsible for the entropy stabilization observed. The stability of the deprotonated complex decreases when the size of the alkaline earth ion increases (log Kf from 4.50 i 0.02 for magnesium ion to 3.67 i 0.02 for barium ion at 0.05 ionic strength), or when the charge of the ion decreases (log Kf of 1.43 i 0.02 for sodium ion at 0.07 ionic strength). The pKa's and magnesium complex f0rmation constants of several PAA analogs: phosphonoformic acid, 2-phosphon0pr0pi0nic acid, and Pierre-Henri C. Heubel 3-phosph0nopropionic acid were determined. The results show evidence for a relationship between the amount of magnesium complexed at pH 7 (physiological pH) and the biochemical activity of the ligand. AUX "QUATRE AUTRES" ET A NANCY 11° ACKNOWLEDGMENTS The author wishes to thank Professor Alexander I. Popov for his counseling and for his guidance into the field of chemical research and scientific writing. Thanks are also extended to Professor Bernard Tremillon for introducing me to the analytical field and for his friendship. Financial aid from the Department of Chemistry, Michigan State University, and the National Science Foundation was appreciated. The help of Frank Bennis and Wayne Burkhardt in keeping the NMR spectrometers in operating condition is acknowledged. Dr. John G. Hoogerheide (also known as Long John, Mutt, the siamese brother, etc.) is thanked for his friendship and constant encouragement during this whole research project. The friendship of Dr. Jenshang Shih is also acknowledged. The author would like to extend his appreciation to Dr. Neil Armstrong f0r being the second reader, to Dr. Andrew Timnick for his willingness to discuss technical problems, and to Dr. Donald Hard for his help in computer programming and usage. The author wishes to thank Nancy A. Fedewa fbr her love, patience and concern throughout this research project. Her everyday smile was greatly appreciated. Finally, the author is deeply grateful to his family. The love, consideration and encouragement of the "four others" contributed to a large extent to the successful completion of this research project. TABLE OF CONTENTS Chapter Page LIST OF TABLES ....................... ix LIST OF FIGURES ....................... xi PART I: MAGNESIUM-25 NMR STUDIES .............. 1 l HISTORICAL REVIEW .................... 2 1.1. Introduction ................... 3 1.2. Magnesium-25 NMR ................. 3 1.3. Proton NMR .................... 3 1.4. Ligands ...................... 6 1.5. Conclusion .................... 9 2 MATERIALS AND METHODS .................. 10 2.1. Materials ..................... 11 2.1.1. Reagents ................. 11 2.1.2. Solvents ................. 13 2.1.3. Molecular Sieves ............. 15 2.2. Instruments .................... 15 2.2.1. Flame Emission .............. 15 2.2.2. Nuclear Magnetic Resonance ........ 15 2.2.2.1. DA-60 NMR Spectrometer . . . . 16 2.2.2.2. Bruker HFX-90 NMR Spectrometer ......... 16 2.2.2.3. Bruker WH-180 Superconducting NMR Spectrometer ....... 16 2.2.2.4. Operating Procedures ..... 17 2.2.3. Infra-red Spectra ............ 18 iv PART II: 1 2.3. 2.4. NMR Sample Preparation .............. Data Handling and Miscellanei ........... RESULTS AND DISCUSSION ................. 3.1. 3.2. 3.3. 3.4. 3.5. 3.6. 3.7. 3.8. 1 .1. 1. 1. \l 0" 01 b (A) N O C O O 0 Introduction ................... Problems Specific to Magnesium-25 NMR ....... Solubilities of Magnesium Salts in Nonaqueous Solvents Study of Magnesium Salts ............. 3.4.1. 3.4.2. Water .................. Nonaqueous Solvents ........... 3.4.2.1. Chemical Shift ........ 3.4.2.2. Line-Width .......... Donor Number Correlation ............. Complexation ................... I.R. Studies ................... Conclusion .................... PHOSPHONOACETIC ACID STUDIES ........... HISTORICAL REVIEW .................... Introduction ................... Phosphonoacetic Acid ............... Analytical Techniques ............... Nuclear Magnetic Resonance Studies ........ Acidities and Actinide Complexes of PAA ...... Industrial Applications .............. Biological Properties of PAA ........... 1.7.1. 1.7.2. Herpesviruses .............. Herpesvirus DNA polymerase ........ 18 19 20 21 21 23 24 24 24 24 33 36 39 42 42 44 45 46 47 49 49 50 52 52 52 53 1.7.3. Herpesvirus Mode of Action ........ 1.7.4. Inhibitory Power of PAA ......... 1.7.5. Drugs Against Herpesviruses ....... 1.7.6. Toxicological Effects of PAA ....... 1.8. Phosphonoacetic Acid Derivatives ......... 1.9. Conclusion .................... MATERIALS AND METHODS .................. 2.1. Reagents ..................... 2.2. Resins ...................... 2.3. Spectroscopy ................... 2.3.1. Visible Spectroscopy ........... 2.3.2. Nuclear Magnetic Resonance Spectroscopy ............... 2.4. Cyclic Voltammetry ................ 2.5. Potentiometry ................... 2.6. Coulometer .................... 2.7. Sample Preparation ................ 2.7.1. Cyclic Voltammetry ............ 2.7.2. Sodium Ion Electrode Measurements 2.7.3. pH Measurements ............. 2.8. Data Handling ................... RESULTS AND DISCUSSION ................. 3.1. Acidity Studies of PAA .............. 3.1.1. Spectroscopic Studies of PAA ....... 3.1.1.1. Phosphorus-31 and Proton NMR . . . . . ......... 3.1.1.2. Carbon-13 NMR ......... vi 53 54 57 57 60 61 62 63 64 64 64 64 65 65 67 7O 7O 7O 70 72 73 74 74 74 81 3.1.2. Potentiometric Studies of the pK's of PAA .................. 87 3.1.2.1. Coulometric Titrations . . . . 87 3.1.2.2. pK's of PAA .......... 90 3.1.2.3. Thermodynamics of Protonation .......... 99 3.2. Complexation Studies of PAA ............ 102 3.2.1. Spectroscopic Studies .......... 102 3.2.1.1. Visible Spectra ........ 102 3.2.1.2. Sodium-23 NMR ......... 103 3.2.1.3. Phosphorus-31 NMR ....... 103 3.2.2. Cyclic Voltammetry ............ 105 3.2.3. Potentiometric Studies of Complexation ............... 111 3.2.3.1. Sodium Ion Electrode Study . . 111 3.2.3.2. Glass Electrode Study ..... 111 3.2.3.3. Thermodynamics of Magnesium Complexation . . . . 120 3.3. Phosphonoacetic Acid Analogs ........... 120 3.3.1. Spectroscopic Studies of PFA ....... 121 3.3.2. Determination of the pK's ........ 121 3.3.3. Complexation ..... ' .......... 122 3.4 Conclusion .................... 125 APPENDICES A SODIUM-23 NMR ...................... 126 A.1. Introduction ................... 127 A.2. Results and Discussion .............. 127 B APPLICATIONS OF THE COMPUTER PROGRAM KINFIT4 ...... 135 vii 8.1. Calibration of a pH Glass Electrode ........ 8.1.1. Program Function ............. B.1.2. Subroutine EQN ............. 8.1.3. Data Sample ............... 8.2. Determination of the Formation Constants of 1:1 and 2:1 Complexes From NMR Data ...... 8.2.1. Program Function ............. B.2.2. Subroutines ............... 8.2.2.1. Subroutines EQN for Data Simulation ........ 8.2.2.2 Subroutine EQN ........ 8.2.3. Data Sample ............... C APPLICATION OF THE COMPUTER PROGRAM MINIQUAD 76A TO THE DETERMINATION OF EQUILIBRIUM CONSTANTS FROM POTENTIOMETRIC DATA ................... C.l. Program MINIQUAD 76A ............... C.2. Data Input Instructions for MINIQUAD 76A ..... C.3. Data Sample .................... BIBLIOGRAPHY ........................ viii 136 136 137 139 139 Table II III IV VI VII VIII IX LIST OF TABLES Page Key Solvent Properties and Correction for Magnetic Susceptibility ............ 14 Line-Width of the Magnesium-25 NMR Signal in Different Solvents ........... 34 Limiting Chemical Shift of Magnesium-25 in Various Solvents ................ 38 Vibrational Frequencies of Free and Complexed Acetonitrile .............. 43 Carbon-13 Chemical Shift of 0.7 M_ PAA y§_pH ..................... 84 Acidity Functions of Citric Acid at Ionic Strength 0.l ................ 91 Acidity Functions of PAA at Various Ionic Strengths at 25°C .............. 97 Thermodynamic Quantities for Acid Association and for the Mg-PA Complex ...................... 101 Half-wave Potential of the Cd(II)/Cd Couple in the Presence of PA y§_Thallium Amalgam ................ 110 ix XI XII XIII PAA Complexes With Metal Ions ....................... Formation Constants for PAA-Magnesium Complexes at Various Ionic Strengths at 25°C ................. Complexation Formation Constants of PAA Analogs at 25°C and 0.05 Ionic Strength .................. Formation Constants of Sodium-(lZ-C-4) Complexes and Their Chemical Shifts ........ 112 115 123 134 Figure LIST OF FIGURES Page Structures of Representative Cryptands and Crowns (Cryptand: The Numbers Represent the Number of Oxygens in Each Bridge, Crown: The First Number is the Number of Atoms in the Cycle, and the Second, the Number of Oxygens ..... 7 Magnesium-25 Chemical Shifts 0f MgBrz in Acetonitrile ................... 26 Magnesium-25 Chemical Shifts of M9012 (A) and MgC1214 (I) in PC ................ 27 Magnesium-25 Chemical Shifts in THF [MgClz (X), MgC1214 (0)] andAcetone [MgC12 (A), MgClZI4 (I)] .................... 28 Magnesium-25 Chemical Shifts of MgCl2 in Methanol (A), DMSO (0) and DMF (I) and 0f MgC1214 in DMF (x) .................. 29 Magnesium-25 Chemical Shift of MgCl2 !§_ Mole Ratio of Iodine in Acetonitrile ......... 31 Magnesium-25 Chemical Shift of MgCl2 y; Mole Ratio of Iodine in PC .............. 32 Limiting Magnesium-25 Chemical Shift in Various Solvents y§_Their Gutmann Donor Number ........ 37 xi 1O 11 12 13 14 15 16 17 18 19 Structures of Various Phosphonic Acids: Phosphonoacetic Acid, Phosphonoformic Acid, 2-Phosph0n0pr0pi0nic Acid, and 3-Phosph0n0propionic Acid ............. DNA Synthesis and Mode of Action of PAA ....................... Block Diagram of the Constant Current Coulometer ................. Phosphorus-31 Chemical Shift of 0.3 M PAA y§_pH ..................... Comparison of the Phosphorus-31 Chemical Shifts and the Species Distribution of PAA y§_pH ..... Carbon-13 Chemical Shift (Referenced to Dioxane, Which is 68.1 ppm Downfield of TMS) of the CH2 Carbon of 0.7 M_PAA y§_pH ............. Change in JC-P Coupling Constant of the CH Carbon of PAA y§_pH ................ Plot of the pKZ of PAA gs Ionic Strength, the Line is a Linear Least Squares Fit ....... Plot of the pK3 of PAA y§_Ionic Strength, the Line is a Linear Least Squares Fit ....... Species Distribution Plot of PAA at 0.05 Ionic Strength ................... Temperature Dependence of 1n Kf for Three PAA Complexes, the Lines are Linear Least Squares Fits . . xii 55 68 95 96 98 100 20 21 22 23 24 25 26 27 28 29 Phosphorus-31 Chemical Shift of 0.3 M_PAA and of a 1:1 Mg:PAA Solution y§_pH ............ 104 Phosphorus-31 Chemical Shift of 0.3 M PAA Xi Mole Ratio of Magnesium at pH Above 9 ........ 106 Plot of E1/2 as a Function of Log [PA] for . the Cd2+-PA System (Reference: Thallium Amalgam) . . . 109 Coulometric Titration Curves of the PAA and PAA-Mg2+ Systems ................... 113 Plot of Log Kf of the Monoprotonated PAA-Mg2+ Complex y§_Ionic Strength, the Line is a Linear Least Squares Fit ............... 116 Plot of the Log Kf 0f the Completely Deprotonated PAA-Mg2+ Complex y§_lonic Strength, the Line is a Linear Least Squares Fit ........ 117 Species Distribution Plot of the PAA-Mg2+ System at 0.05 Ionic Strength ................ 119 Species Distribution Plot of the PFA-M92+ System at 0.05 Ionic Strength ................ 124 Sodium-23 Chemical Shift y§_M01e Ratio of 12-C-4. NaBPh4: (O): 0.1 M in PC, (I): 0.034 M in THF: (0): 0.04 M NaI in Methanol: (D): 0.04 M NaClO4 in Water ................... 129 Sodium-23 Chemical Shift y§_Mole Ratio of 12-C-4. NaBPh4: (I): 0.04 M in Pyridine, 0.1 M in DMSO (A). Acetone (x). DMF (A), Acetonitrile (o) ..... 130 xiii PART I MAGNESIUM-25 NMR STUDIES CHAPTER 1 HISTORICAL REVIEW 1.1. INTRODUCTION The main purpose of this thesis was to study the complexes formed between magnesium ion and phosphonoacetic acid (PAA) by different physicochemical techniques. Techniques such as visible UV spectroscopy are not very useful to study the environment of group I or group II metal ions in their stable oxidation states. In the case of most of the alkali metal ions, the observation of the metal nuclide resonance and the study of both the line-width and the chemical shift of the signal gave interesting information about the chemical environment and the complexation of these nuclei in their first oxidation state (1). For these reasons, the magnesium-25 nuclear magnetic resonance (NMR) technique was applied to the magnesium-phosphonoacetic acid system. 1.2. MAGNESIUM-25 NMR However, the properties of the magnesium-25 nucleus are not very favorable for NMR studies. The magnesium-25 nucleus has a low nuclear magnetic sensitivity (the senstivity per nucleus at constant field is only 2.68 x 10'2 of the sensitivity of the proton), a low natural abundance (10.05%), a low resonance frequency and a quadrupolar momentum (the nuclear spin is 5/2). In 1951, Alder and Yu (2) located the resonance frequency of magnesium-25 in a 4.6 M solution of magnesium chloride in water near a frequency of 3.9 MHz, in a magnetic field of 11 k6. They also measured the magnetic moment of magnesium-25 by NMR. However, because of the poor properties of the magnesium-25 nucleus mentioned earlier, few studies of magnesium-25 NMR have been performed. 4 Using continuous wave technique, Ellenberger and Villemin (3) reported magnesium-25 NMR studies of aqueous solutions. They found lineridths of 40 to 45 Hz for the different solutions used. However, within the precision of the measurements, the resonance frequency seemed to be independent of the salt concentration. Subsequently, Dougangt 91. (4) and Dickson and Seymour (5) determined the Knight shift of magnesium-25 in magnesium metal, that is the paramagnetic shift of the nuclear resonance due to the delocalized conduction electrons of a metal (named after W. D. Knight who was the first one to observe it in copper in 1949). Bryant (6) used magnesium-25 NMR to study aqueous solutions and reported a line- width (full width at half height of the peak) of 3.8 i 0.5 Hz for a 1.5 M_MgC12 solution, which is an order of magnitude smaller than the one reported by Ellenberger and Villemin (3). Bryant's use of magnesium-25 NMR was essentially to determine the chemical rate 0f exchange (based on measurements of the line-width) of magnesium between complexed and uncomplexed sites of ATP type phosphate solutions. In the absence of chemical shift change in aqueous magnesium solutions, Magnusson and 80thner-By (7)used the change in line- width gs pH of acidic ligand solutions (e.g., citric acid, ATP) to approximate the formation constants of the corresponding complexes. Although this method seems unsound because of the various factors affecting the lineawidth of the signal and because of the precision of such measurements, their values are within one order of magnitude of other published values. The above authors reported a variation of the line-width from 5 to 9 Hz when the concentration of magnesium ions was increased from 1.0 to 4.0 M, They also reported that above pH 7, a solution of magnesium ions and EDTA gave one broad peak, thus showing that the exchange was fast compared to the NMR time scale. The widest line-width reported for magnesium-25 NMR was in the case of complex formation with ATP (1300 Hz). Simeral and Maciel (8) are among the last ones so far to have reported magnesium-25 NMR studies, and those were also done in aqueous solutions. After having taken the viscosities of the different solutions into account, the line-width data have been discussed in terms of four models for quadrupolar nuclear spin relaxation in the systems studied. The line-widths reported vary from 3 to 11 Hz, but once again, no chemical shift change either with the concentration or the counterion was observed. 1.3. PROTON NMR Proton magnetic resonance has been used to study the solvation number of magnesium ion in acetone-water and methanol-water mixtures (9-15). The inner solvation shell consisted of six water molecules as long as the mole ratio of water to magnesium was equal to or higher than 6. Matwiyoff and Taube (10) also determined that in methanol-acetone mixtures, methanol was a much better solvating agent than acetone. In all those solutions the total solvation number was 6. The preparation of anhydrous magnesium salts presents a difficult problem. With one exception [16, redox reaction between magnesium and mercury (II) halide in tetrahydrofuran], all the purifications or synthesis (17) of magnesium chloride or bromide 6 found in the literature involve the fusion of the halide and its separation from magnesium oxides by filtration in a stream of dry nitrogen in a quartz tube (the melting points of MgC12 and MgBr2 are 708° and 700°C, respectively). The affinity of magnesium ions for water is such that direct heating of the hydrated salt under active vacuum leads to the formation of magnesium oxides. Thus, according to Halliwell and Nyberg (18), the hydration enthalpy of magnesium is -459 kcal/mole, respectively four, five, six and seven times the ones of lithium (~124 kcal/mole), sodium (-97 kcal/mole), potassium (-77 kcal/mole) and cesium (-63 kcal/mole). Since all of the above magnesium-25 studies were restricted to aqueous solutions, it was of interest to us to study this resonance in nonaqueous solvents. 1.4. LIGANDS Cyclic polyethers, or crown ethers, were developed by Pedersen (19 and references therein). They have a remarkable ability to form stable two-dimensional complexes with alkali metal cations. Struc- tural formulae of some typical crown ethers are shown in Figure 1. Formation of such complexes distorts the spherically symmetrical electric field around the solvated metal ion which may result in a broadening and/or a chemical shift change of the metal NMR signal. There are (19-23) several reports of syntheses, spectroscopic and crystallographic studies of the ligand lZ-crown-4 (12-C-4), also known as the ethylene oxide cyclic tetramer (1,4,7,10 tetraoxa- cyclododecane). The ligahd 12-C-4 is a liquid at room temperature with a specific gravity of 1.104, a boiling point of 238°C at 760 mm .Amcmmaxo mo .8952 m5 .283 mg..— vce .298 05 5 mac; Co c3532 65 3 $452 amt... 0.: "586 6375 53 E. Swabs .8 $4.52 9.» 23¢..qu €2.52 2:. 653.28 mczocu E; mucougcu 2333333. “B 83323 .p 953... moo 5% cup—305-2 ticgu au— 3 0 .CL. 8 of mercury and a vapor pressure of 0.03 mm at room temperature (24). The cavity size of this crown is 1.2 to 1.5 A, which is approximately the size of the magnesium ion (1.3 A diameter from crystallographic measurements). In a study of the complexation of crown ethers, Chock (25) found that the enthalpy of hydration involved in the desolvation step and the binding energy for a given ligand were the most important factors in the mechanism of complexation and the stability of the resulting complex. These observations, added to the great affinity of magnesium ion for water, probably explain why the crystal structure of MgClz, 6H20-12-C-4 complex (20) consists of octahedral Mg(H20)§+ units very similar to those in MgClz, 6H20, located on crystallographic axes, with 8 of the 12 water hydrogens forming hydrogen bonds to six different chloride ions, which are located in tetrahedral holes, while the remaining four hydrogens of the Mg(H20)g+ unit form hydrogen bonds to ether oxygens from four different cyclomer molecules. Leong gt_gl, (24, 26) reported some toxicological studies of 12-C-4. Rats exposed to the vapors of 12-C-4 for six hours daily for six to eight days exhibited variable degrees of anorexia, loss of body weight, asthenia, hindquarter incoordination, testicular atrophy, auditory hypersensibility, tremors, convulsions, moribond conditions and death, depending on the exposure concentration. Cryptands are diaza-polyoxamacrocycles. These ligands with bridging nitrogens were first synthesized by Lehn £3 21, (27, 28). They can form very stable complexes with alkali and alkaline-earth cations. The length and number of ether bridges may be changed to vary the size of the inner cavity of the cryptand. The term cryptand 9 refers to the ligand and cryptate to the complex. The stability of the cryptate complexes depends to a large extent on the size relationship between the three-dimensional cavity of the cryptand and the diameter of the complexed ion. In general, cryptands form much stronger complexes with the alkali metal ions than the crown ethers, provided that ligands with approximately the same cavity size are compared. Recent reviews (29-33) contain extensive compilations of complexation stability constants. Cryptand 0211 is shown in Figure 1 (21] refers to the number of oxygens in each hydrocarbon chain). Its cavity size is approximately 1.6 A. Very little work has been done on the complexation of Mg2+ ion by C211. Lehn and Sauvage (29) reported log Kf for a 1:1 complex to be 2.5 i 0.3 and 4.0 i 0.8 in water and in 95/5 methanol/water mixture respectively, while Morf and Simon (30) found a log Kf value of less than 2 in water. 1.5. CONCLUSION Very little work has been done concerning the NMR properties of magnesium-25. In order to use this technique to study the magnesium-PAA complexes, it is necessary to perform preliminary studies of the influence of the environment of the magnesium ion upon its NMR properties. CHAPTER 2 MATERIALS AND METHODS 10 11 2.1. MATERIALS 2.1.1. REAGENTS - The following chemicals were all reagent grade and used without further purification. The code is: Aldrich (A), Alfa inorganics (AI), J. T. Baker (8), Drake brothers (0), Fisher scientific company (F), Mallinckrodt (M), Matheson Colleman and Bell (MCB) and G. F. Smith Company (S): Barium chloride (M), barium oxide (MCB,F), cadmium nitrate tetrahydrate (8), citric acid monohydrate (F), cobalt chloride (AI), cobalt sulfate (8), copper sulfate (AI), dioxane (F), ethylene diammine tetraacetic acid, disodium salt (MCB), glacial acetic acid (M), hydrochloric acid (M), lithium bromide (MCB), magnesium acetate tetrahydrate (8), magnesium bromide hexahydrate (F), magnesium chloride hexahydrate (MCB), Magnesium perchlorate (S), magnesium sulfate (M), manganese chloride (AI), nickel bromide (AI), potassium hydroxide (MCB), potassium iodide (F), potassium permanganate (B), 3(trimethylsilyl)-1-propanesulfonic acid, sodium salt hydrate (known as 055, the water soluble TMS: (CH3)3Si(CH2)3SO3Na, nHZO, is an internal standard for use with 020) (A), sodium carbonate (F), sodium hydroxide (0), sodium iodide (F), tetraphenylarsonium chloride (S) and zinc chloride (M). Tetramethyl ammonium hydroxide was from either Eastman organics or MCB and was obtained as a 10% solution in water. Phosphonoformic acid, trisodium salt hexahydrate and 3—phosphonopropionic acid were kindly furnished by J. Reno from Dr. J. Boezi's group (Biochemistry Department) and used as received. Cryptand 211 was from E. M. Lab, and 12-crown-4 from Aldrich; 3-phosphonopropionic acid was from Richmond Organics. 12 Several indicators have been used: bromocresol green and eriochrome black T (F), phenol red (A), and phenolphthalein (F). The following chemicals were dried in the vacuum oven, unless otherwise Specified (Company, time, temperature): magnesium metal (8, 6h., 60°C), sodium tetraphenylborate (8, 48h., 25°C), lithium perchlorate (F, 96h., 190°C, regular oven), mercuric chloride, bromide and nitrate (M, 6h., 60°C), sodium chloride (M, 24h., 150°C), sodium perchlorate (S, 24h., 110°C) and mercuric thiocyanate (Ventron Alfa Products, 6h., 60°C). A flame and K. Fischer analysis of 12-crown-4 showed that the mole ratios of Na+ ion and water to 12-crown-4 were 0.26% and 2%, respectively. No traces of lithium or potassium were detected. Iodine (Baker Analyzed reagents, ACS or USP grade) was purified by double sublimation: 10 g of iodine are first sublimed with 5 g of potassium iodide (in order to eliminate the iodine chloride and iodine bromide) at 70-80°C. It takes about 10 hours to sublime 10 g of iodine. Then the obtained iodine is sublimed with 4.3 g of granulated barium oxide (in order to remove the water) under the same conditions. Bromine (Dow Chemical) was dried over P205 for 24 hours and then distilled under atm05pheric pressure. Magnesium salts: anhydrous magnesium bromide was first prepared by direct reaction of bromine over magnesium in ether, but contamina- tion of the solution by bromine made the solution useless for NMR purposes. Therefore, magnesium chloride and bromide were prepared by using the method described by Ashby and Arnott (10). A redox reaction between a mercury (11) halide and an excess of magnesium metal in a tetrahydrofuran (THF) diethylether mixture l3 (kept under reflux for 4 to 6 hours) leads to the formation of the corresponding magnesium salt and a magnesium amalgam. It was found here that THF alone gave better results for M9012 and MgBrz than the THF-ethylether mixture. Mercury (11) bromide (86 g., 0.24 mole) and metallic magnesium turnings (6 g., 0.25 mole) were refluxing in 1.4 1. of THF for 6 hours (in the case of magnesium chloride, the concentrations were doubled: 0.5 mole in 1.4 1.). The solution then was transferred to a dry box and the amalgam and excess of magnesium were removed by filtration through filter paper (because of the presence of liquid mercury and of clogging problems, a glass frit could not be used). Tetrahydrofuran is then evaporated on a rotavap under vacuum and the salt is dried under vacuum at 80°C overnight. Omission of this step lead to supersolubility problems, due to the contamination of the solution by the small amounts of THF remaining in the salt. The salt was stored in a dry box where the nonaqueous solutions were prepared. Attempts to prepare Mg(N03)2, Mg(SCN)2 or Mg(BPh4)2 by the same technique failed, as well as attempts to prepare Mg(8Ph4)2 by either using the method given in the literature for the lithium salt (34), or the potassium salt (35). This latter method was used to prepare Hg(8Ph4)2. 2.l.2. SOLVENTS - The solvents were obtained from the following companies: Acetone (F), acetonitrile (MCB), dimethylformamide (DMF, F), dimethylsulfbxide (DMSO, F), methanol (F), propylene carbonate (PC, A), tetrahydrofuran (THF, M). (See code in Part II.2.1.l.). Useful solvent parameters are listed in Table I. 14 ooo.o ooo.o o.mm em.mn seam: mee.o+ e-.o- o.o~ mm.“ Razhv catacosuzeespah omm.o+ owp.o- _.mp o.mm Auav eueeoeeeu eeeFAaoee mmm.o+ m~3.o- n.m~ A.Nm Pegasus: ~m3.o+ Fem.o- w.m~ «8.83 Aomzov eewxoc_=mpxeoeswo mpm.o+ mom.o- c.m~ p~.om Aazav eUPEessoc_»eoeewo own.o+ omm.o- _.5_ m.~m ep_sopeoueo< ao.p+ mem.o- o.AF A.o~ eeooae< 253 8 I. :53 8-5 89.52 8:8 28.28 85; Sm co cowuomccou co cowuumccou m.c:aEp=o owcpuopmwc .quFWaPpnmumsm uwumcmoz com cowuuoccou use mowpcmaocm uco>Pom xox .H mpnm» 15 All solvents except methanol were dried over calcium hydride or activated 3A and 4A molecular sieves (3A has a cavity size that just fits the water molecule) for 5 hours, followed by a distillation at atmospheric pressure (except for DMSO for which a static low pressure was applied). Usually 10 g of CaHZ or 100 ml of Sieves were used for one liter of solvent. Methanol was dried by double distillation over magnesium (36). The water contents of the solvents used were measured with the help of a Karl Fischer automatic titrator and were found to be always less than 50 ppm's. 2.l.3. MOLECULAR SIEVES - Molecular sieves 3A (Davison Chemical Company, W. R. Grace and Co.) and 4A (MCB) were activated by drying at a maximum of 400°C for 6 hours under nitrogen which was dried by passing it through concentrated sulfuric acid. 2.2. INSTRUMENTS 2.2.1. FLAME EMISSION - Flame emission experiments were performed using a Heath EU-703-70 flame unit, an EU 700 scanning monochromator and EU-70l-30 photomultiplier module containing an R-44-UR Hammamatsu photomultiplier tube. Output from the P.M. tube was recorded through a Heath EU-703-3l photometric readout amplifier. The total consumption burner used hydrogen at 1 psi and air at 15 psi. Samples were aSpirated for at least 30 seconds with distilled water blanks between samples. Duplicate runs were made for each standard and unknown. 2.2.2. NUCLEAR MAGNETIC RESONANCE - The NMR spectra were collected on three different instruments, all operated in the Fourier transform mode. 16 2.2.2.1. DA-6O NMR Spectrometer - A modified Varian DA-60 NMR spectrometer was used for magnesium-25 and sodium-23 NMR measurements. The instrument consists of the magnet of a Varian DA-60 Spectrometer equipped with a wide band probe capable of multi- nuclear operation, as described by Traficante gt_al, (137). The field is at 14.09 kG and the frequency of the RF unit is 56.44 MHz. A detailed description of the spectrometer has been given elsewhere (133), The resonance frequency of the sample can be obtained by changing the synthesizer frequency, instead of changing the magnetic field. Before each experiment, the probe and image filter were tuned to the resonance frequency of the nucleus under consideration (3.672 MHz f0r magnesium-25 and 15.87 MHz for sodium-23). The field is locked by a home-built lock probe on the proton resonance of the upfield sideband of H20). Pulse generation and data collection were performed with the help of a Nicolet 1083 computer with 12K of memory. The tubes were 25 and 15 mm 00 (magnesium-25 NMR) and 10 mm 00 (sodium-23 NMR). 2.2.2.2. Bruker HFX-90 NMR Spectrometer - The phosphorus- 31 and sodium-23 NMR spectra were obtained by using a Bruker HFX-90 NMR spectrometer equipped with a Nicolet 1083 computer with 12K of memory, a Diablo disc memory unit and a Nicolet 293 I/O controller. The spectra were obtained at a frequency of 36.44 MHz (phosphorus-31) and 23.81 MHz (sodium-23). Spinning tubes of 10 mm 00 were used. 2.2.2.3. Bruker WH-180 SuperconductinngMR Spectrometer - This instrument was used to collect magnesium-25 NMR spectra at 11.01 MHz, and proton spectra at 180 MHz. The field strength is 17 42.28 kG. An external 020 lock was used. Due to the superconducting magnet, the field was very stable, and results were obtained without using an internal lock for the nonaqueous samples. Solutions were run in 20 mm 00 nonspinning tubes, the minimum amount of solution for each tube is 8 m1. Preliminary studies showed that due to the broadening of the peak by quadrupolar relaxation, the absence of spinning did not affect the line width to a measurable extent. 2.2.2.4. Operating Procedures - The chemical shifts were measured against external references. A 4.0 M solution of aqueous NaCl was used for sodium-23, and a 4.0 M solution of aqueous MgClz (DA-60) or a 4.0 M 75/25 HZO/DZO MgClz solution (8-180) for magnesium-25. AS shown by Maciel (8), there is no difference in chemical shift for these last two solutions. The phosphorus-31 NMR chemical shifts are reported relative to 85% phosphoric acid. According to the convention used in this thesis, a positive change in chemical shift of sodium-23, magnesium-25 and phosphorus-31 NMR signals corresponds to a shift to higher field strength than that of the reference signal. All chemical shifts are corrected for the bulk diamagnetic susceptibility difference between water and the pure nonaqueous solvent. For each solution, the contribution of the salt to the susceptibility of the solution is assumed to be negligible: a reasonable assumption as shown by Templeman and Van Geet (37). The formulae given by Live and Chan (38) for normal (equation 1) and superconducting (equation 2) magnet were used: 18 acorr = 6obs -%m(x€ef _ Xsample) (1) 5corr = aobs +%fl(xcef _ xsample) (2) In the case of magnesium-25 and phosphorus-31 NMR, when the top of the peak was too flat to allow precise determination of the chemical shift change, the precise frequency of the peak was obtained by averaging the two frequencies at half height. 2.2.3. INFRA-RED SPECTRA - Infra-red spectra covering a range of 4000-600 cm'1 were collected on two Perkin-Elmer grating IR spectrophotometers: models 237 B and 457. The frequencies were calibrated using polystyrene reference peaks. Solutions and nujol mulls of solid samples were placed between sodium chloride mull plates. 2.3. NMR SAMPLE PREPARATION Nonaqueous solutions of magnesium salts were prepared in a dry box. When the dissolution of the magnesium salt lead to the formation of a light precipitate [presumably Mg(OH)2], the solution was filtered in the dry box. Since magnesium has a large affinity for water, the water content for each sample was measured after each NMR measurement, and data points were kept only if the mole ratio of water to magnesium was less than 20%. In these conditions, the lowest concentration studied in nonaqueous solvents was 10"2 M, which corresponds to a water content for the solution of 36 ppm. In most cases, the lowest concentration was 3 x 10"2 M. The magnesium concentration in the different solvents was measured 19 by diluting a small sample (usually 1 ml) of solution with distilled water followed by the usual titration by EDTA in NH3/NH4Cl buffered medium. Since 12-crown-4 and cryptand 211 are liquid at room temperature, the preparation of the solutions containing them was completed outside the dry box. The amounts of the above ligands or sodium salts used were determined by weight. Suitable diultions were performed using 2, 5, 10, 25 and 50 ml volumetric flasks. 2.4. DATA HANDLING AND MISCELLANEI Extensive use was made of the CDC 6500 computer with several Fortran IV programs (see Appendices B and C). Elemental analyses were done by Chemalytics Inc. of Tempe, Arizona, and/or by Spang Microanalytical Laboratory, Eagle Arbor, Michigan; the latter giving more precise results. Melting points were determined on the Fisher-Johns melting point apparatus. The instrument was calibrated from 40°C to 300°C by means of melting point standards (Hoover). Analyses for water were accomplished with a Photovolt Aquatest II automatic Karl Fischer titration apparatus. CHAPTER 3 RESULTS AND DISCUSSION 20 21 3.1. INTRODUCTION Previous studies in this laboratory and elsewhere have Shown that the NMR of alkali metal ions offers a sensitive Probe of the environment of these ions in different solvents and solvent mixtures. The purpose of this study is to determine whether magnesium-25 NMR would be a suitable technique to study the influence of the cation on the ionic equilibria and ionic species present in various solvents. The exchange of ions between different environments is usually rapid with respect to the NMR time scale, resulting in only one resonance signal at an average frequency determined by the magnetic shielding and nucleus population in each of the sites. Alteration of parameters such as concentration, counterions and solvent produces changes in the relative proportion and type of environment which may be reflected by a change in chemical shift and/or line-width of the observed resonance. 3.2. PROBLEMS SPECIFIC TO MAGNESIUM-25 NMR In the first place, magnesium ion has a very large affinity for water. Since most organic solvents cannot be obtained completely anhydrous, it is obvious that in such cases meaningful chemical shifts for magnesium can only be obtained if the concentration of water is much smaller than the concentration of the salt. We found that in the present study a mole ratio of water to magnesium of 20% was the upper water content limit. For example, a solution of magnesium chloride in acetonitrile, with a mole ratio of water to magnesium of 50% gave rise to a change in chemical shift of 2 ppm, compared to a solution with a mole ratio of 10%. 22 Secondly, prior to this study no one had performed any magnesium-25 NMR studies of nonaqueous solutions of magnesium salts, and, therefore, one of the problems was to determine whether the change in chemical Shift would be large enough to be detected easily with our instruments (after all, lithium ion which has approximately the same size as magnesium ion has a chemical shift range of only a few ppm). Thirdly, several instruments have been used in this study, with a large increase in sensitivity for each change of instrument: thus, with the first insert and probe, the sensitivity of the DA-60 was such that the magnetic field did not need to and could not be tuned: in these conditions a 0.05 M aqueous solution of magnesium chloride required 8000 scans (50 minutes of time with 4K of memory and 5000 Hz sweepwidth for the computer) for a signal to noise ratio of 2.5 (the signal to noise ratio is defined as the ratio of the signal over the largest noise, multiplied by 2.4). The first improvement was to use a better probe, which allowed to study the same solution in much better conditions (6000 scans gave a signal to noise ratio of 8). The second and best improvement was to use the Bruker 180 Spectrometer with which a 0.046 M aqueous solution of MgClZ can be analyzed in 10 minutes. Finally, another major problem specific to magnesium is the following - either the ion is strongly solvated and does not form contact ion pairs, resulting in no chemical shift change with salt con- centration, and small line-width change (due to the change in viscosity of the solution), or there is ion pairing, which, if it results in a change in chemical shift, also significantly increases the line-width. 23 For this latter reason and because of the low sensitivity of the instrument, only few solutions could be studied with the DA-60 (mainly aqueous, methanol and acetonitrile solutions). 3.3. SOLUBILITIES OF MAGNESIUM SALTS IN NONAQUEOUS SOLVENTS The solubilities of MgCl2 and MgBr2 in nonaqueous solvents are fairly small. The best cases studied are represented by acetonitrile (2.0 M), acetone (1.2 M), methanol (1.0 M) and dimethylformamide (DMF, 0.9 MD. For propylene carbonate (PC) and tetrahydrofuran (THF), the solubility is approximately 0.3 M, while it decreases to 0.1 M for dimethylsulfoxide (DMSO), and to 10'3 M for pyridine and nitromethane. It was first thought that the solubility in nitromethane was around 1.0 M, but it was found later on that THF impurity in the salt (from its synthesis) was responsible for this high value. At the time the salt was only dried at room temperature under vacuum for 24 hours. Upon dissolution in a given solvent, the THF molecules solvate the cation thus increasing the "solubility" of the salt. On occasion, after the preparation of a 1.0 M solution of MgClz, the salt would precipitate out, leading to a final solubility of the salt of less than 10"2 M, The solubility of magnesium acetate (Mg(0Ac)2) was found much smaller as a rule than the solubility of MgClz or MgBrZ, since except in methanol (0.1 M) all the other solvents tested (aminoethanol, formic acid, f0rmamide, DMF, ethylenediamine, DMSO, THF, nitromethane, pyridine, acetone) dissolved at most 10"2 M_Mg(OAc)2, which is not suitable for a magnesium-25 NMR study in the present conditions. 24 3.4. STUDY OF MAGNESIUM SALTS 3.4.1. MAI§R_- Five different magnesium salts have been studied in water, Mg(OAc)2, MgClz, MgIZ, Mg(NO3)2, and Mg(ClO4)2. No change in chemical shift was observed when the concentration of each of these salts was changed. In the case of four of these salts [MgClzz 4.6 to 0.1 M, MgIz: 1.8 to 0.1 M, Mg(NO3)2: 3.0 to 0.05 M and Mg(C104)2: 3.2 to 0.05 M) the change in lineewidth was small: from 17 to 7 Hz, while for Mg(OAc)2 the line-width changed from 43 to 8 Hz for a corresponding change in concentration of 2.0 to 0.24 M, However, in the latter case, the concentrated solution was much more viscous than the other salts solutions, and this fact is probably responsible for the broadening of the peak. These results are similar to the ones reported by Maciel gt_al, (8). As expected, magnesium ion is strongly solvated by water, and from these measurements, it appears that if ion pairs are formed between the different ions in solution, these are at least solvent shared ion pairs. 3.4.2. NONAQUEOUS SOLVENTS 3.4.2.1. Chemical Shift - It has been previously observed (39, 40) that the contact ion pair equilibria strongly depend on the donor ability of the solvent molecule as well as on the bulk dielectric constant of the medium. Although PC and acetonitrile have high dielectric constants (65 and 37.5, respectively), their donor abilities are low on Gutmann's scale (41) (15.1 and 14.1, respectively). Gutmann defined a donor ability scale for solvents as the negative 25 AH-value in kcal/mole for the interaction of the solvent with SbCl5 in a highly diluted solution of dichloroethane. We see from Figures 2 and 3 that the magnesium-25 chemical shifts of MgBrz and MgClz solutions are concentration dependent and, therefore, that there is contact ion pair formation. These results are in agreement with the data obtained by sodium-23 and lithium-7 NMR in PC and acetonitrile solutions where the chemical Shifts of NaI and NaSCN or LiBr and LiI are strongly concentration dependent (42, 43). Tetrahydrofuran and acetone are interesting solvents in that they have low dielectric constants (7.58 and 20.7, respectively), but reasonable donor abilities (20.0 and 17.0). As can be seen from Figure 4, there is contact ion pair formation even between Mg2+ and the Cllé ions. 0n the other hand, solvents like methanol, DMSO or DMF that have both high dielectric constants and high donor numbers and are good solvating agents show no or very little ion pairing (see Figure 5). For the methanol and DMSO solutions, the chemical shift does not change with concentration (1.1 and 3.5 ppm, respectively), while for the DMF solutions, the chemical shift changes from -4 to 4 ppm for a concentration change of 0.9 to 0.03 M. We mentioned above that we attribute the concentration dependence of magnesium-25 chemical Shifts to the formation of contact ion pairs, i.e., to cases where the anion directly replaces a solvent molecule or molecules in the inner solvation shell of the cation. It seems reasonable to assume that gross variations in the chemical shift and/or the line-width of the magnesium-25 26 .e_peoweeueu< :0 News: ca mocpem _eupeeeu m~-e=mme=mez .N ee=m_a 2.58.84 5 «cm as. 0.. 0.0 _ 40m: (wdd) ””9 26 rung ow 628.2362 5 N28: .3 3:5 .8255 3-5856: .N 2:2... 0.. 0.0 o__.=_c2oo< c. N6 02 0 Own 27 rm 0.): . no .8 E i exam: 2; 3 ~53. .8 3:5 8:85 3.552%: .m 23: ONO H 0.0 0N: .28 .2! :38: . 9 . mmmcmez e we: .0 N u< use :3 53%: .00 ~82: “.E 5 3:5 59:55 mm :5 0.0 a 0 0:30 3 5 E 0.0 0.0... ”to; 6.. s _ 29 be as: o._ .8 “=3 5 3.6.sz co 2; i “=3 nee i 85 .3 .828: 5 N89,. co Stem .6385 8.53886: .m 28.: 00 0 4 _ 0N... 40. 30 NMR signal are related to a change of its direct chemical environment, that is,it reflects the influence of its nearest neighbors. It is well known that NMR gives information about the inner solvation shell of the nucleus under consideration. Replacement of the solvent molecule(s) by the anion may either increase or decrease the electron density of the cation. Replacement of solvent molecules by halide ions increases the electron density around the magnesium ion resulting in a downfield chemical shift of the Mg-25 resonance. Similar behavior has been previously observed with Li, Na, K, and Cs salts studied by Li-7, Na-23, K-39 and Cs-l33 NMR (44). Another evidence f0r the formation of contact ion-pair in the above solutions is the result obtained by studying the ngz-I2 mixtures in different solvents. It is well known (45) that halogen molecules interact with halide ions to form trihalide ions which 7 are quite stable in nonaqueous solvents (e.g., Kf of 106°6, 10 and 1010 for 13, Br§ and C13 ions respectively in acetonitrile). We can expect, therefore, the IZCl‘ complexes to be stable, and, in the concentration range used (0.7 to 0.05 M), we can assume that at a 1:1 mole ratio of iodine:halide ion, all the halide ions will be complexed. The size of the anion will become significantly larger and thus will have a much smaller tendency to form ion-pairs with magnesium ions. The effect of iodine upon Mg2 +-X' ion-pairing is shown in Figures 6 and 7 - there is a linear change in chemical shift [similar to the one reported by Erlich and Popov (46) in the case of sodium-23 NMR] y§_mole ratio of iodine to Mg2+ ion up to a value of 2 where all the halide ions are complexed. The largest 31 62.5.5884 .... 9:8. ca 6.5.2. 6.6... m.» ~82. .6 at...“ .8265 3.5.528... ... 6.53. 0 _. oz_\_~__ 0.~ 0.. b . . Rod ... N. 362 o._.=_co.oo< c. «.5 as. 0N. .iom 32 .um cw mcmuom mo ovumm upoz.mm N—umz mo ukwgm pmuwemso mmusswmmcmoz .m mgampm _.~§_\_~._ oe o.~ o _ _ 8.. o_- mmnoé as: m: I o .8 c. «a 22 Lo. lulllfllln‘ o C «En—3 scum 33 effect of ion-pairing was observed in the case of PC solutions, where the change in chemical shift due to ion-pairing was more than 30 ppm. For solvents with a medium value of the dielectric constant (PC, acetonitrile, DMF), the interaction of the magnesium ion with the halide ion is decreased to such an extent that a magnesium-25 NMR concentration study of the complexed salt results in a constant chemical shift. ' 0n the other hand, for solvents with a low dielectric constant (THF, acetone), the interaction between the two ions is decreased but is still noticeable, as shown by Figure 4. The poor solvating ability of acetone, and of THF had been shown by sodium-23 and lithium-7 NMR (47, 48). 3.4.2.2. Line-Width - Another interesting feature of the magnesium-25 NMR signal is the line-width. In the case of strong solvating agent (i.e., with both high dielectric Constant and high donor number), the line-width of the signal is small. It is approximately l0 Hz in water, methanol and DMSO solutions. In DMF, concentration change from 0.03 to 0.9 fl_increases the line-width from 30 to 127 Hz. As can be seen from Table II, the line-widths in other solvents are large, which also implies contact ion-pair formation. Indeed, although PC solutions are more viscous than aqueous solutions, this is not the case for acetone solutions, and, therefore, line-broadening of the signal must be due to other effects. It is well known that the line-width of an NMR signal is related to the relaxation time T; of the nucleus by the equation: 34 Table 11. Line Width of the Magnesium-25 NMR Signal in Different Solvents Magnesium Line Width Solvent Concentration (m) at Half Height (Hz) Methanol 0.768 12 0.0543 10 THF 0.325 235 0.1248 220 0.0325a 146 PC 0.369 460 0.111 390 0.111a 134 DMF 0.913 127 0.0319 30 0.0319a 15 DMSO 0.1227 37 0.01227 11 Acetone 1.221 3l0 0.1195 215 0.1196a 150 Acetonitrile 2.045 62 0.423 27 0.6303 24 aWith a mole ratio of iodine to magnesium equal to or higher than 2. 35 _ 1 Av (3) 1/2 nTE with 1 = 1 + YAHO * (4) T2 T2 2 AH where g represents the contribution to the line-width due to inhomogeneity of the magnetic field AHO, v is the magnetogyric ratio, a constant for a given nucleus, and T2 is the spin-spin relaxation time. Since the observed relaxation time is assumed to be in the motionally narrowed limit so that T1 = T2, with T1 being the spin lattice relaxation time, we can relate the line-width of the signal to the five categories of spin lattice relaxation mechanisms: dipole-dipole relaxation, chemical shift anisotropy, scalar coupling, spin rotation and quadrupole relaxation. We can then express the observed results as follows: 'I = 1 " = . . 1" _TT 'fifg 1, exp Rl, d1p-d1p Rl , chem shift an + R + R + R 5 l, scal l, spin rot l, quad ( ) It has been shown (49) that in the case of quadrupolar nuclide resonance, the line-width is generally determined by the quadrupolar interaction, the contributions from the other relaxation mechanisms being negligible. The interaction Hamiltonian for the quadrupole relaxation has the form: Hq = 1.0.1 and represents the interaction between the nuclear spin (I>l/2) and the electric field gradient at the nucleus. As reorientation occurs, the components of the quadrupole coupling tensor Q become random functions of time and provide a relaxation mechanism. We then have: 36 R=R= 3 21+3, n2 92 2 (6) 1 2 40 X 1?}21-1)\1 +7) (—hg'g) TC where n is the asymmetry parameter and (equ)/h is the quadrupole coupling constant, I is the nuclear spin of the observed species and Tc is the translational correlation time. Thus, the line-width of the NMR signal depends on three terms: n, Tc and the quadrupole coupling constant. When ion-pairs are formed, the asymmetry parameter and the field gradient (and hence the quadrupole coupling constant) will change, and we will observe an increase in line-width. In the absence of ion-pairing, the line- width will be affected by the solvent cage. In a strongly solvating medium, the solvent molecules are held tightly around the metal ion, and the change in electric field around the nucleus is minimal, therefore, the line-width of the signal is small. 0n the other hand, when the magnesium ion is surrounded by a less solvating agent, like acetone or PC, the solvent molecules around the magnesium are held less tightly, and having more latitude to move, cause an increase in the inhomogeneity of the field, thus allowing the magnesium ion to relax faster, which in turn is reflected by a wider line-width. 3.5. DONOR NUMBER CORRELATION It has been shown previously (50, SI) that in the case of the sodium-23 and potassium-39 resonances there is a roughly linear relationship between the magnitude of the chemical shift in a given solvent and the donicity of the latter expressed by the Gutmann donor number. Similar plot is shown in Figure 8 for the magnesium-25 resonance, and the data are given in Table III. 37 .20 06 ...mnsaz cocoa 5555 ..EE. 1m.» 3539.... maotm> ..., p.25 $3.55 3:533:32 9.3.2.3 .m 939... Om ON 0. _ 4 ”=9: _ O?! l O .4025...sz OM20 O NZOPNUQ 39.". l m m w l O. L V— BEEP—MS 38 Table III. Limiting Chemical Shift of Magnesium-25 in Various Solvents. Solvent 5 (ppm)a Donor Number Dé:;::§:1c Acetone 3.0 17.0 20.7 Acetonitrile 14.0 14.1 37.5 DMF 5.5 26.6 36.71 DMSO 3.5 29.8 46.68 Methanol 1.1 25.7 32.7 PC 13.5 15.1 65.0 THF -4.0 20.0 7.58 Water 0.0 33.0 78.9 a!§_4.0 M_aqueous susceptibility. MgClz as external standard. Corrected for bulk 39 It is readily seen that for a majority of solvents there is a definite correlation between the magnitude of the downfield chemical shift and the donor number. Five of the eight solvents studied seem to fall on a reasonable straight line, but the correlation is not good in the case of THF, acetone and methanol. For the latter, a similar behavior was observed in the case of cesium-133 and potassium-39 resonances. 3.6. COMPLEXATION Only preliminary studies have been performed in this area. An aqueous solution of phosphonoacetate and 0.05 M_magnesium chloride at pH above 8 (so that all the PAA is under the deprotonated form) gave the following results: for a mole ratio of 0, 0.05, 0.1 and 0.2, the line-width changed from 7 to 16, 25 and 50 Hz, respectively, and no change in the chemical shift was observed. For higher mole ratios, the line-width was too large and the signal could not be detected within a reasonable amount of time. It is seen from the above results that we have a fast exchange, compared to the NMR time scale, between the free and the complexed magnesium ion, therefore, the line-width and the chemical shift are the average of the free and complexed magnesium signals. If we define Ta and Tb as the lifetime of the magnesium nucleus in sites A and B, u and “b as their resonance frequencies and 'l' as: T = T 1: (Ta + Tb) (7) we will have two resonance signals if 1:) 2 nIva-vb) (3) otherwise, we will have the population average resonance. 40 Likewise, the line-width of the complexed magnesium ion is large, due to quadrupolar broadening since the ligand, being unsymmetrical, produces a nonuniform electrical field around the nucleus. Similar studies performed with EDTA at pH 9-10 showed a somewhat different result, the line-width of the peak stays constant when the mole ratio ligand to metal increases from zero to close to one. At the same time, the signal becomes weaker and weaker, and finally for a mole ratio of one or above, it disappears. These results show that we have a slow exchange, compared to the NMR time scale, of free and complexed magnesium. The line-width of the complexed magnesium signal is again much too large to allow its detection. These results are in contradiction with the ones reported by Magnusson and Bothner-By (7). Since the formation constant of the MgZ+-C21l complex in water and in 95/5 methanol/water mixture was reported as 102'5 and 104'0, respectively (29), and since C211 and Li+ ion form a complex with an exchange between the free and complexed Li+ ion that is slow compared to the lithium-7 NMR time scale, the complexation between C211 and MgClZ has been studied in methanol. Kauffmann (52) from carbon-13 MIR studies of 0.1 [1 solutions of C211 reported the formation of strong complex between magnesium and C211 in 95/5 methanol/water mixture. She found that in the 25°C-85°C temperature interval, the spectra of a 2:1 mixture of C211 and MgClz contained two series of signals, one corresponding to the free ligand, while the other one was characteristic of the cryptate. 41 From equation (8), it can be seen that if the rate of exchange (r = 1/1) between the free and complexed magnesium is slow compared to the C-13 NMR time scale, such is not necessarily the case for the magnesium-25 NMR time scale, since magnesium-25 resonates at a lower frequency and has a smaller chemical shift range than carbon-13. Indeed, in the present case, the exchange between the two sites is fast since only one signal is observed. The line width increases from 10 to 22 Hz for the solutions of 5 x 10'2 11 MgC12 with a mole ratio of ligand to magnesium of O and 1, respectively. The change in chemical shift, however, was small - from 1.1 to 0.7 ppm and, therefore, could be attributed to a random experimental error. As an aside, the perchloric acid salt of C211 is insoluble in methanol. This property could be used to purify C211 or to recover it. Another complexation study with lZ-crown-4.(12-C-4, which has a cavity size comparable to the size of the magnesium ion) in DMF has been performed. In this case, no change in either line-width nor in chemical shift was observed. Either there is no complexation between Mgz+ ion and 12-C-4, or both free and complexed ions have the same chemical shift. The first hypothesis seems more reasonable in view of the solvating properties of DMF (e = 36.71, donor number = 26.6). On the other hand, at this early stage of development of magnesium-25 NMR in nonaqueous solutions, only qualitative results can be obtained as far as complexation is concerned, since there is no quantitative results concerning the ion-pairing in these solutions. 42 3.7. I.R. STUDIES Some I.R. studies of 2.0 M_solutions of MgBr2 in acetonitrile showed a behavior of the solvent similar to the one observed by Janz et al.(53)by Raman spectroscopy in the case of silver ion. The splitting of the solvent bands corresponding to the C-C and CEN symmetric stretches (a]) for complexed solvent molecules is shown in Table IV. The splitting is due to the presence in solution of both free and complexed acetonitrile. 3.8. CONCLUSION We determined in this study that the chemical shift range of the magnesium-25 NMR signal is larger than 30 ppm, and that the NMR of magnesium-25 can be a useful tool to study solvent effect and ion-pairing of magnesium salt in nonaqueous solutions. The line broadening and chemical shift change have been explained by quadrupolar relaxation in terms of ion-pairing and solvent- magnesium interaction. 43 Aceveepom Ncmmz.m Nm~.~v _-eU OF-mom~ .ommm .oemm _-EU mmm .epm .owm upmcumeepee< eexepeEOU 2-58 omNN .oeNN F-ea m_m epwcuweeeee< mesa eeeetem ewcpeseam zmu sebacem uwceeeexm 8-8 eeweapom .mpwepwcoumu< nmxmpasou uco amen we mmwocmzamcm pmcomuwenw> .>H mpnmh PART II PHOSPHONOACETIC ACID STUDIES 44 CHAPTER 1 HISTORICAL REVIEW 45 46 1.1. INTRODUCTION Just as the use of antibiotics revolutionized the treatment of bacterial disease in the 1940's, new antiviral agents promise to revolutionize the treatment of viral diseases in the 1970's and 1980's. For the first time, man will be able to combat viral infections, halting the replication of viruses and preventing their spread. The advent of this era may well be one of the most important milestones in the continuing battle against infectious diseases. The genesis of the antiviral era has been quite different from that of the antibiotic era. The latter was launched by the discovery of penicillin, a relatively broad spectrum antibiotic whose value was readily apparent. The antiviral era is being launched with a number of narrow-spectrum agents whose value has been more difficult to establish. Antibiotics were introduced at a time when only a limited amount of testing was necessary to introduce a new drug to the market and when the need to treat wounded soldiers during a major war greatly accelerated that testing. Antiviral agents are being introduced at a time when consumer safety is the paramount concern, necessitating a great deal of expensive, time-consuming clinical testing. The development and testing of antibiotics was largely subsidized by the federal government because of the national emergency, and the first antibiotics were quite profitable for the companies that put them on the market. The development of antiviral agents, in contrast, has been financed by the drug companies with 47 only limited support from the government, and no company has made a profit on one. In some ways, it is remarkable that the antiviral agents have been developed as rapidly as they have been. 1.2. PHOSPHONOACETIC ACID Herpes viruses are a large family of deoxiribonucleic acid (DNA) viruses that are the causative factor of diseases, such as cold sores of the mouth, genital lesions, eye infections, varicella (chicken pox) and shingles. By some estimates as many as 10% of all Americans over the age of 18 have recurrent Herpes infections three or more times per year, and more than 70% of Americans are thought to have antibodies in their blood that indicate a prior Herpes infection. Some types of persistent Herpes infections are believed to be associated with initiation of cancer. Few antiviral drugs exist against Herpesviruses, and most of them are nucleoside analogs, and, therefore, they are potential teratogenic agents (i.e., they could cause birth defects if injected during pregnancy). Phosphonoacetic acid (PAA), which is not a nucleoside analog, and has been found active against quite a few different Herpesviruses is, therefore, a very promising compound. Phosphonoacetic acid (Chemical Abstract #4408-78-0) is an organophosphorus compound (see Figure 9), it has an inhibitory effect on the replication of Herpesviruses, and the mode of inhibition is thought to involve complexation with magnesium ion. 48 .umu< umcowaogqocosamogalm ucm .ku< upcownogqococqmogmim .uwo< owELomococamosg .uwo< owuwomococamosa IO IO 1 _ suumxuumxo .. 4.0: o = O 0.04 U_Zo.n_0mn_OZOIn_mOIn_lm IO IO _ mu-zo-e-o: mu m” _ = 180 0.04 U_ZO_QomQOZOIQmOIQ1N ”mcmu< upcosamosm maowgm> mo mmgzuuzgum .m wczawu 10 ...o, _ .Ahvnu nuiunvv+ o __ o «6.1.x: 0.04 0.2m0u0201amOIa IO IO, N . _ So... IUunTOI . 2. A43": 904 Ochoa. OZOIQmOIa 49 This triacid, as well as several of its salts (calcium, barium, copper, manganese, zinc and silver), were first synthesized in 1924 by Nylen (54). Several syntheses have been proposed since then (55). Phosphonoacetic acid is a triacid; it is a white solid, with a molecular weight of 140.03, and has a sharp melting point at l42-l43°C; it will also be referred to as phosphonoacetate (PA), its completely deprotonated salt. 1.3. ANALYTICAL TECHNIQUES Several infra-red (IR) studies of PAA or its derivatives have been published in the literature (56-59). Its trimethylsilyl derivative has been studied by gas chromatography-mass spectroscopy and by gas chromatography (60); a single peak is obtained when examined in polar (0V-17) and nonpolar (SE-30) gas chromatographic phases. The above study is of importance for the biochemist, and the technique has been used to monitor the amount of PAA in blood. Phosphonic acid derivatives have also been separated by thin layer chromatography (61 ) . It has been found (62) that phosphonates can inhibit the transfer of copper ions or c0pper complexed ions to a mercury electrode because of their surface active properties. Investigations have been conducted on the adsorption of PA on hematite, silica and graphitized carbon (63, 64). 1.4. NUCLEAR MAGNETIC RESONANCE STUDIES Nuclear magnetic resonance studies involving the measurement of the carbon-l3 or phosphorus-31 chemical shift or of the coupling 50 constants between C and P or P and H have been conducted on PAA or its ethylester. For carbon-l3 NMR studies, the acid was dissolved in methanol (65) and the chemical shifts vs TMS are 35.8 and 168.6 ppm, for the CH2 and the COOH carbons, respectively, while the coupling constant between C and P is 132.0 Hz. The ester has been studied as a neat liquid (66). Some phosphorus-31 NMR studies have been performed on PA (67). The solution was basic (pH=l4) with tetramethylammonium hydroxide (TMAH). The chemical shift reported !§_85% phosphoric acid (contained in a capillary tube inside the sample tube) is -l3.l ppm, no indication being given whether it is a downfield or an upfield shift. Riess §t_al, (67) found that a several fold change in either phosphorus or alkalinity concentrations caused less than 0.2 ppm change in the chemical shift. No mention was made by them of a triplet signal due to the coupling between the P and the CH2 hydrogens. Complexes between lithium and two PAA esters have been studied in tetrahydrofuran (THF) (68). 1.5. ACIDITIES AND ACTINIDE COMPLEXES OF PAA Aside from these industrial uses, all the other studies made so far were concerned with either the extraction of rare earth elements, or biological applications. It is interesting to note that most of the studies concerning the biological applications of PAA have been published in the last five years, and that they represent approximately 80% of all the studies ever performed on PAA. 51 Organophosphorus compounds have been found to be good extractants of rare earth elements (69), and more specifically, Elesin _$“El- (70, 71) investigated the complexation of PAA with americium, curium and promethium ions, as well as determined the acidity dissociation constants (pK‘s) of PAA. Several comments can be made about this work: the dissociation constants of PAA were determined against a background of ammonium perchlorate at an ionic strength I = 0.2, although ammonium perchlorate is known to be a weak acid. Moreover, sodium hydroxide was used as a titrant in this research, and since it has been shown (72, 73) that phosphate ions form weak complexes with alkali metal cations, the values of the pK's obtained (1.14, 4.33, and 7.31 at ionic strength 0.2, which, extrapolated to zero ionic strength gives 1.37, 4.84, and 8.50) are probably too small. The investigation of the complex formation of Am3+, Cm3+, and Pm3+ was conducted by the method of ion exchange on the cation exchange resin KU-Z, also against a background of ammonium perchlorate, at 0.2 ionic strength. Only protonated complexes were studied, and the log Kf values extrapolated to zero ionic strength are 2.75, 5.15 and 3.3 for the M(H2L)2+, M(HL)+ and M(HL)§ complexes, respectively (LH3 represents the freePAA). The formation constants were found to be equal for the three cations. Mao £3 11. (74) reported pK's of PAA of 2.6, 5.0 and 8.2 for the carboxylic and two phosphono groups, respectively. Their correlation between pK's and acidity functions is most probably 52 incorrect (see p 74), besides, the conditions under which those values have been determined are not specified. 1.6. INDUSTRIAL APPLICATIONS The practical applications of PAA or its esters are numerous, and several patents mention it for widely different purposes: it has been used as a flame retardant in thermoplastics (75), as a corrosion inhibitor for iron-containing water conducting systems (76), to prevent the formation of spots in photographic materials due to the presence of particles of heavy metals or their derivatives (e.g., iron and rust) (77), as an insecticide (78), as an herbicide (79), or a plant growth regulator (80). Phosphonoacetic acid has also been proposed as a treatment against warts (81), or Herpesviruses (82, 83). 1.7. BIOLOGICAL PROPERTIES OF PAA The popularity of PAA in the biochemical field lies in its ability to inhibit the replication of Herpesviruses. Three good reviews have been, or are going to be published on that matter (84-86). 1.7.1. HERPESVIRUSES - Herpesviruses (Herpesvirus in the U.S., Herpes virus in England) is the name of a genus of viruses that cause diseases affecting both man and animal. Some of these diseases are fatal, and it has been suggested recently that Herpesvirus may be directly or indirectly associated with cancer in several animals. 53 Herpesviruses have also been indirectly associated with naso- pharyngeal and cervical carcinomas in humans [carcinoma: solid tumor derived from epithelial tissues,is the major form (85%) of human cancer]. Other types of Herpesviruses are: Equine Herpesvirus, Varicella-zoster virus (causes chicken pox, shingles), Marek‘s disease Herpesvirus (the main cause of death among chicken in commercial poultry, and the reason of the creation of the Poultry Research Laboratory at M.S.U.), Epstein-Barr virus (responsible for infectious mononucleosis), Cytomegalovirus (among the diseases: pneumonitis, Cytomegalovirus mononucleosis), Herpes Simplex virus [causing cutaneous lesions (Herpes Dermatitis), mouth sores, cornea infection (Herpes Keratitis: responsible for an estimated 18,000 cases of blindness each year in the United States), venereal diseases]. The list is far from being exhaustive. 1.7.2. HERPESVIRUS DNA POLYMERASE - It has been found for several types of Herpesviruses that the virus induces its own DNA polymerase (DNA polymerase is an enzyme involved in the production of DNA). Three general classes of DNA polymerases are found in cells of all vertebrates: DNA polymerase a, B, and y. The greek letters designate their order of discovery: 0 and B are the most important ones. Herpesvirus induced DNA polymerases have been reported for several Herpesviruses. 1.7.3. HERPESVIRUS MODE OF ACTION - The mode of action of Herpesvirus is as fOllows: the virus enclosed in a capsid protein 54 comes in contact with a cell, at this point the protein opens and the viral DNA penetrates into the cell, it then travels through the cell to the nucleus, enters it, and uses the nucleus system to produce ribonucleic acid typical of the virus (RNAHV). The RNAHV leaves the nucleus, and following the usual pattern, forms a DNA polymerase typical of the virus. This DNA polymerase goes back to the nucleus where it creates its own DNA (DNAHV); from this point on, the loop is closed, the production of DNAHV increases a lot, and within a few hours or a few days, depending on the virus, the cell dies and many new viruses are then released. 1.7.4. INHIBITORY POWER OF PAA - Leinbach gt_gl, (87) proposed two solutions to explain the inhibitory power of PAA on the Herpesvirus of turkey induced DNA polymerase. The basic pathway used by the enzyme to build up the Herpesvirus specific DNA is shown in Figure 10. In the first step, the DNAHv binds to the enzyme, in the second step deoxynucleoside triphosphate (dNTP) binds to another site of the enzyme, upon which the length of the viral DNAHV molecule is increased and inorganic pyrophosphate (PPi) is formed. In a third step, PP, is released and in the final step DNAHV is released. The loop is closed, and the procedure can start again. Actually, in each cycle, the DNAHV molecule is not exactly increased by one link, but rather the two complementary strands of DNA are separated and each forms the template for synthesis of a complementary daughter strand. From the results of their kinetic studies, Leinbach gt_gl, (87) proposed the mechanism presented in Figure 10. In the presence of 55 .42.. ..e 5:2 .3 so: 2; 205.5 55 .2 23: 11......m ...-22¢ a: ll m epz. . . .=:385 1 1 1 m2: boarL 78853: 5.238 - 250m 088 E38 1 .sz53 mmqm m2: _ H mofimmzmo .35ch _1rm» <43 .8 «.22 em .mp 323322 77 in w.» < 44¢ m0“. w>m30 ZO_._.Dm_m._.m_n_ mwawam 3:0. 0 wk 0 n v _ _ _ ON CV 00 0m 00. o\o WE “.0 «=22 mm m. 3 Egg 0. m_ 78 artificially chosen coordinate notation which allows the chemical shift expression to be stated in a particularly simple functional form, using a minimum number of hybridization parameters. This derivation has been applied to several forms of molecules, including molecules of the MPZZT type. According to this derivation, the chemical shift may be treated as the summation of o-bond and n-bond contributions. The distinction made between a and n bonding is that 0 molecular orbitals involve only §_and p_orbitals while g orbitals alone are involved in the n bonding. Moedritzer (119) further developed the quantum derivation of Letcher and van Watzer, and applied it to the case of the chemical shift change of a series of oxyacids of phosphorus .22 pH. Thus, for molecules in which the phosphorus has a coordination number of 4, the phosphorus-31 chemical shift is comprised of a 0- bond contribution determined solely by the prorbital occupation and a n-bond contribution corresponding to g_orbitals only. The chemical shift can be expressed as: 6 = 60 + 60 + 6Tr (9) where 6 is the chemical shift referenced to 85% H3P04, 60 is the absolute chemical shift of the reference standard and 60 and 61r are the o and 0 contributions, respectively. The theory showed that 61r is negative and is proportional to the increase in total occupation of the d1r orbitals of the phosphorous, being independent of the distribution of this total character among the various bond. For P atoms having 4 neighbors, 61r = -l47nfl, where n1T is the total number of electrons in the dTr orbitals of 79 the phosphorous atom being viewed by phosphorus—31 NMR. On the other hand, the prorbital occupation was treated in terms of several parameters: the angular geometry, consisting of the molecular symmetry and bond angles, and the polarity of the bonds. Moedritzer then proposed that the change in phosphorus-31 chemical shift could be expressed as the sum of small variations in: l) the effective electronegativity of the neighboring oxygens due to association of hydrogen, 2) the bond angles and 3) the total occupation of the g_orbitals. Therefore, according to Moedritzer, equation (9) becomes: A6 = CAXO - 147An1r - Aae (10) where 4X0 is the change in the effective electronegativity of the oxygens caused by their association with the hydrogen(s), n1T is the change in the total number of electrons in the d1r orbitals of the phosphorous induced by variations in the character of the P-O bond brought about by hydrogen association, A0 is the increase in 0P0 bond angle caused by association of the hydrogen, and A and C are numbers which may vary from one oxyacid to another. In the case of PAA we can write equation (10) slightly differently: ‘ A6 = CAxo + BAxR ~147Ann -DAn: - A00 (11) where B and D are constants, AXR reflects the change in electro- negativity of the organic group (acetic acid group presently), R n of n electrons, and the other terms have been defined previously. An corresponds to the change in P-O n bond from rearrangement 80 Upon protonation of the phosphoric acid moieties, we have an increase in electronegativity of the corresponding oxygen atom, corresponding to a positive contribution to the chemical shift. On the other hand, we also have feedback of n electrons from the P-O bond onto the phosphorus resulting in an increase in n electrons and a negative contribution to the chemical shift. The change in bond angle has not been considered here and is probably not completely negligible although of minor contribution. According to the theory developed by Moedritzer, in the case of phosphonic acids, we must have a larger contribution to the chemical shift from the feedback mechanism than from the change in polarity (while the opposite is true f0r the H3PO case). No 4 quantitative evaluation of each term has been done so far, and the above argumentation explains why the above theory of phosphorus— 31 NMR chemical shift has been referred to as a "tug of war" (118), but the theory fits our experimental results. Studies performed by Riess gt_gl, (67) showed that although R the terms involving changes in n1T or 0 are not negligible, they are much smaller than the contribution of the electronegativity of the carboxyl group. Furthermore, no n bonding is involved in the P-CH2 bond, and the effect of the rearrangement of the n electrons around the P-O n bonds due to change in electron distribu- tion of the CH2 group is most probably negligible. We can also assume that the change in A6 (O-P-O angle) upon protonation of the acetic acid moiety which is at the other end of the molecule is negligible. 81 By bonding a proton to the carboxylic group, the electronega- tivity of the related oxygen increases, thus increasing the electro- negativity of the two carbons [corresponding to a decrease in prelectrons, as shown by 13C NMR (see below)] which corresponds to a positive contribution to the chemical shift as shown in equation (11). This formalism also applies to the replacement of the acetic acid moiety by the formic acid moiety as shown in Part II.3.3.1. No conclusion can be drawn from the change of the coupling constant JP-H y§_pH because of the precision of the measurements when JP-H is determined from 31F MR spectra. Whatever is the change, it will be small anyway since JP-H changes between 19.5 and 21.5 Hz with a precision of':0.5 Hz. The value is as expected (118) for the P-H coupling through an intermediate nucleus, i.e., between 15 and 25 Hz. On the other hand, the line-width of the central peak changes from 3 Hz at pH 1.0 to 8 Hz at pH 6.0 and back down to 3 Hz at pH 9.0. A study of the chemical shift of phosphorus-31 in the free acid y§_concentration shows an upfield shift from -16.4 ppm at 1.0 M_to 2 -13.3 ppm at 10' M. If we again take a pK1 of 1.5, a simple calculation shows that a 1.0 M solution of PAA contains 84% of triprotonated acid, while a 10'2 M solution contains 80% of diprotonated acid. According to Figure 13, the chemical shift should then be below -16 ppm, and at -13.3 ppm for the l M and 10'2 M solutions, respectively, in good agreement with the experimental results. 3.1.1.2. Carbon-13 NMR - 0f more interest for the pK-acidity function correlation is the study of the change in 82 carbon-13 NMR chemical shift and JP-C coupling constant v_s_ pH. Generally speaking, an upfield shift for any NMR measurement can be related to an increase in §_(diamagnetic) electron density or a decrease in p_(paramagnetic) electron density. As a rule, shifts in carbon-13 NMR are mostly related to §_electron density. However, the PAA case is slightly different, and resembles the studies performed on the deprotonation of carboxylic acids. Upon deprotonation, the carbon-13 NMR chemical shift of the carboxylic carbon moves downfield by approximately 5 ppm, due to a decrease of the electronegativity of the oxygen, which corresponds to an increase in p_electron density around the oxygen atom and, therefore, around the carbon. The decoupled carbon-l3 NMR spectrum of 0.4 M aqueous PAA shows two doublets, corresponding to the COOH and the 9H2 carbons (at 172.8 and 173.1 ppm, and 34.4 and 40.7 ppm, respectively at pH 1.0). The doublets correspond to the P-C coupling, with a larger coupling constant for the carbon closer to the phosphorus. Figure 14 shows the change in chemical shift !§_pH for one of the peaks of the CH2 carbon signal, and the values for the four peaks are listed in Table V. As expected, we have a downfield chemical shift. When the pH is increased, however, the change starts with the removal of the phosphoric acid proton, and the overall change corresponding to the removal of the three protons is 7.5 ppm for the COOH carbon signals, and 6.2 and 5.5 ppm for the CH2 carbon peaks. In this case, again, the pH was adjusted with tetramethylammonium hydroxide. 83 .3... .3» s: m we .8 .523 9...... 2: ee Amze ee upe_eezeo see ..me 3, £3.33 .eeexe_o e3 emceeceeeav Sewem .eu_eeeu m_-eeeccu .3. eceu.e In 0_ m. m .h m n 3» mm m“ _ 0 .VM1 1. NMI Emu - w 1 ON! 4®N1 Io _ Ia .m> 681901410: ...o «:22 on. o Tab1e V. 84 Carbon-13 Chemica1 Shift of 0.7 fl PAA !§_pH pH -£H2 0.76 173.0 172.7 40.4 34.2 1.48 173.7 173.4 40.9 34.8 2.44 174.4 174.0 41.3 35.4 3.51 174.7 174.5 41.6 35.7 4.46 175.4 175.1 42.1 36.2 5.47 176.8 176.5 43.1 37.2 6.45 177.6 177.3 43.7 37.7 6.93 177.7 177.4 43.8 37.9 8.11 178.3 178.1 44.3 38.4 8.95 179.4 179.1 45.2 39.4 85 On the other hand, if we consider the JC-P coupling constant for the ng carbon (see Figure 15), we have a decrease in JC-P from 125.5 to 118.3 Hz for a change in pH from 0.8 to 4.0 followed by a slight increase from 118.3 to 119.5 Hz (pH changes from 4.0 to 6.5)and a final decrease from 119.5 to 112.2 Hz when the pH is brought to 10.6. The JC-H coupling constant has been related by quantum mechanical calculation to the §_e1ectron density around the carbon and the hydrogen (120), an increase in the electronic density corresponding to an increase in JC-H‘ From this assumption, several authors (121, 122) related the JP-C value to the §_electron density around the carbon. 0n the other hand, Tebby (123) reported that the effects of e1ectronegative substituents on coupling are different for phosphorous and carbon. Either they do not increase the §7character of the remaining bonds, or another factor such as n-bonding is introduced. Because of the numerous variables and unknowns concerning the chemical shift and specially the JC-P coupling constant (very little work has been published on that matter and the theory is still missing at the present time, presumably because of the lack of data), it seems unrealistic to try to explain the shape of the JP-C !§_pH curve quantitatively. Qualitatively, it seems reasonable to relate the two decreases in the JP-C coupling constant to the deprotonation of the phosphoric acid moieties. In conclusion, we have correlated the second pK of PAA to the acetic acid moiety in the molecule. The change in direction of the phosphorus-31 NMR signal shift has been explained in terms of changes in the g_electron density around the phosphorus and of changes in electronegativity of the different substituents. .=a_m».< :00??? o _ no: ...o “To: 0 mm. _ ~V__ mw__ mw__ /. afiumw. mme_ .umw_ .mwmw_ ANIV n-0n .m. ae=m_a 87 3.1.2. POTENTIOMETRIC STUDIES OF THEApK'S 0F PAA - Once the pK's of PAA assigned to the different acidity functions, it was of interest to determine precisely the pK's of PAA at low ionic strength and to get the thermodynamic values by extra- polation to zero ionic strength. It was decided to determine these values from measurements performed with a pH electrode, however, because of the reported complexation (72, 73) between phosphate ions of various types and the sodium ions, it was decided to use a much bulkier cation for the base and the supporting electrolyte, i.e., the tetramethylammonium ion. Some titrations have been performed with tetramethyl- ammonium hydroxide as a base, but since the tetramethylammonium hydroxide commercially available is an approximately 1 fl_solution, problems occurred due to the presence of carbonate ions in the base which interfered with the determination of the pK's of PAA (002 cannot be removed in a solution with appH above 7.5 by degassing). Moreover, the commercial product from Eastman contained a precipitate soluble in acid. Attempts to synthesize the base in the laboratory by ion exchange resins were not successful. The use of a cation exchange resin resulted in incomplete exchange of the sodium ion, while the use of an anion exchange resin resulted in only partial exchange of the bromide ion. For these reasons, it was decided to generate the base electrically with a home-built coulometer. 3.1.2.1. Coulometric Titrations - The advantages of that method are the elimination of the preparation and handling 88 of a multiplicity of standard solutions, the increase in the precision in the amount of titrant used, and in the purity of the titrant since the solvent is used to generate it. The reactions involved in the electrode processes are: 2Ho++2e' Z H+2H0 (12) 3 2 2 . _ + _ cathodic 2 H20 + 2 e + H2 + 2 0H reactions (13) 4 0H" Z 02 + 2 H20 + 4 e' . (14) + + _ anodic 6 H20 + 02 + 4 H30 + 4 e reactions (15) + - + 02 + 2 H30 + 2 e + H202 + 2 H20 (16) - -) .- 02+2H20+2e + H202+20H (17) + - -> H202 + 2H30 + 2 e + 4 H20 (18) H202 + 2 e" Z 2 0H" (19) As can be seen from the last four equations, oxygen must be removed from the solutions to prevent formation of H202 and unwanted cathodic reactions. In order to keep the solutions free from oxygen, they were kept under nitrogen atmosphere as explained in the experimental part. In nonbasic conditions, the anodic reaction will generate protons. Since the mobility of proton through the membrane separating the anodic and cathodic compartments is much larger than the mobility of the tetramethylammonium ion, this would result in pollution of the solution by an acidic impurity. To remedy that problem as well as to decrease the ohmic resistance of the system, concentrated (0.2 to 0.5 fl) solutions of the supporting electrolyte tetramethylammonium bromide were used in the anodic compartment. The anodic reactions then become: 89 Br' .2 l/2 Br2 + e" (20) In these conditions, the increase in negative charges due to the formation of base in the cathodic compartment is counterbalanced by a flow of tetramethylammonium ions through the membrane (the membrane is impermeable to the solvent). As mentioned previously (see Part II.2.2.7.3.), a titration of 7 x 10'4 fl_HCl at the working ionic strength is first performed. The EMF of the cell: Glass electrode//Acid solution, 7 x 10'4 [1 at/Thallium amalgam constant ionic strength reference electrode °. -031 4' _R_T. 15. E - E + F 1n [H ] + Ej + Easym + F 1n yH+ (21) where E° is the standard potential referred to the thallium amalgam reference, R, T and F have their usual meaning, E is the junction 3 potential, E is the asymmetry potential and yH+ and [H+] are the asym activity coefficient and the concentration of the hydrogen ion. At a constant ionic strength, the Ej and Easym potentials as well as the activity coefficient of the hydrogen ion are assumed to be constant. Therefore, the EMF of the cell can be expressed as: E = E°' + El x 2.303 log [11*] (22) Using the computer program PHCALIB-KINFIT4 (see Appendix B), the experimental points are fitted to the equation: E = intercept + slope x log ([HT]0 - t : l + residual) (23) F o where [H+]o is the initial concentration of HCl (7 x 10"4 M), i0 is the intensity of the current used to generate the base in mA, t is the time during which the base is generated in seconds, and v0 is the volume of acid in ml. Residual is a concentration 90 of acid or base impurity that could be present in the solution. Initially, the three terms intercept, slope and residual were refined, but it was found that the last two terms are coupled. Since the temperature of the solution is constant to 0.l°C, and pH glass electrodes are known to have nernstian response, the slope term was calculated from the solution temperature and kept constant, and only the two other terms were refined. Once the calibration of the electrode is performed, and in order to keep the junction and asymmetry potentials as constant as possible, the acid to be titrated is added without removing the electrode from the solution. Using this procedure, citric acid was titrated at an ionic strength of 0.1 by tetramethylammonium bromide (TMAB). The pK's obtained are shown in Table VI and compared to literature values. As can be seen the agreement is quite good when compared to the tetramethylammonium chloride supporting electrolyte case (the largest difference between the two sets of values is 0.11 pK units and the smallest 0.01). 3.1.2.2. (pK's of PAA - The pK's of PAA were then determined at different ionic strength. The data reduction was performed by using the computer program MINIQUAD 76A (see Appendix C for an explanation of the data handling) and the three following expressions of the K's: K1 = [Lug] [H+]/[LH3] (24) K2 = [LH=] [H+]/[LH'] (25) 91 Table VI. Acidity Functions of Citric Acid at Ionic Strength 0.l. Temperature Medium pK] pK2 pK3 Reference 20 KCl 2.96 4.39 5.67 a 20 KC1 3.08 4.39 5.49 b 25 KN03 2.79 4.30 5.65 c 25 KN03 2.92 4.39 5.72 d 25 KN03 3.04 4.47 5.80 d 20 NaC104 2.87 4.35 5.68 e 20 NaC104 2.96 4.38 5.68 f 30 NaN03 2.94 4.44 5.82 g 25 (Me)4NC1 2.88 4.36 5.84 h 25 (Me)4NBr 2.89 4.32 5.78 this work aA. Okac and Z. Kolarik, Collect. Czech. Chem. Commun., 24,1 (1959) dT. N. Briggs and J. E. Stuehr, Anal. K. S. Rajan and A. E. Martell, K. K. Tripathy and R. K. Patnaik, J. Indian Chem. Soc., 43, 772 (1966) Inorg. Chem., 4, 462 (1965) eE. Campi, G. Ostacoli, M. Meirone, and G. Saini, J. Chem. , 26, 553 (1964) f C. F. Timberlake, J. Chem. Soc., 5078 (1964) Chem., 47, 1916 (1975) Inorg. Nuclear 9R. C. Warner and I. Weber, J. Am. Chem. Soc., 75, 5086 (1953) hS. S. Tate, A. K. Grzybowski, and S. P. Datta, J. Chem. Soc., 3905 (1965) 92 = [L3'1 [H*1/[LH=1 (26) The relationship between the concentration and thermodynamic formation constant is for K3: 3- t_ KCYL YH K3 27 K311;FF_—— ( ) by using the Guntelberg modification of the Debye-Huckel equation: log yn = - A zfi «T7(1 + 7T) (28) with _ 1.823 x 106 (29) - (cT)3/2 we can rewrite equation (28) as c t 2 2 _ 2 pK3 pK3 A (ZL + 2H ZLH) 777(1 + «7) pkg - pkg 6A «T7(1 + «T1 (30) Similarly we have: pK; = pK: 4A 7T7(1 + «T1 (31) pK: A /T/(1 + If) (32) C PK] Half-way between the second and third equivalence points we have: [TMAE] = 2.5 [PAA] (33) with [TMAE]: concentration of tetramethylammonium ion corresponding to the formation of the equivalent amount of base, on the other hand: [LH‘] = [L3'] = l/2[PAA] (34) therefore, we have: = [S.E.] + 1/2 ([LHE] + 22 [LH=] + 32 [L3']) or I z [S.E.] + 4.5 [PAA] (35) 93 where [S.E.] is the concentration of supporting electrolyte. Similarly, at half-way between the first and second equivalence points: I = [S.E.] + 2 [PAA] (36) With a concentration of PAA: 7 x 10’4 M and a concentration of supporting electrolyte of 0.02, the change in ionic strength during the titration of the third acidity function corresponds to an increase of the ionic strength of 15%. This fact explains why the lowest ionic strength used was 0.02. On the other hand, at an ionic strength of 0.1, a l ppt acid or base impurity in the supporting electrolyte represents an error of 1.5% when referred to PAA. This factor itself limited the acid molarity to 7 x 10'4 minimum. Several comments can be made about the use of equation (31). First, it has been shown that the validity domain of the Guntelberg equation extends to an ionic strength of approximately 0.1, and, therefore, the pK's of PAA were determined at ionic strength lower than that value. Second, when the Guntelberg equation is used, it is equivalent to the Debye-Huckel equation with a diameter of the solvated ion of 3.05 A (at 25°C). Bates (124) proposed to use a value of 9 R for the proton, but we have different ions which sizes are not directly available, and, therefore, the use of the Guntelberg equation seems to be a reasonable approach to the problem. The computer program MINIQUAD 76A reduces the data by calculating the overall formation constants of the equilibria involved, and first starts by refining the pK value, then the pK2K3 and finally 3 the pK1K2K3, therefbre, the error in the value of this last term 94 is a cumulation of the error on each term. For this reason and because the first acidity of PAA is fairly strong (pK1 2 1.5), the extrapolation of the concentration values of pK1 to zero ionic strength was not possible. The thermodynamic values of pK2 and pK3 are 5.11 t 0.04 and 8.69 i 0.05, respectively. The slope of the ionic strength plot is -0.942 and -2.98, respectively, while the theoretical values are -2.04 and -3.05 at 25°C. The thermodynamic plots and the corresponding values are shown in Figures 16 and 17, and Table VII. A non-negligible factor in these potentiometric measurements is the fact that the glass electrode aged. At the end of a two- month utilization period, pK values at different ionic strengths started to fall consistently below the straight lines of Figures 17 and 18 by 0.05 pK units. At the same time, the readings were becoming less stable (precision of i 0.2 or 0.3 mV instead of 0.1 mV). Upon replacement of the combined glass electrode by a new one, the pK values fell back on the line, and no more unstability problems were encountered. The EMF's recorded in two different buffers (acidic and basic) were 165.1/181.4 and -174.6/—155.6 mV for the new/old electrode. The calibration of the electrode is done in the acidic medium and the difference between the two electrode potentials increases from 16.3 to 19.0 mV from acidic to basic pH, or a change of 2.7 mV or 0.046 pK units, which corresponds to the pK difference experimentally observed between the two sets of data. A species distribution plot at ionic strength 0.05 and 25°C is shown in Figure 18. 95 3:... «33cm 33.. .8ch a mm «5.. m5 .5933 FEB in.» 53 mo NV... 93 mo aoE .2 3:3... ONO 0.0 9.0 00.0 0 _ _ memo: , 3o.» 1 on main. mxa Kn . N 8 H3 «5 5.5:? <41 3 v. 96 .....E 8:53 33.. ..8:5 a mm 2.5 23 .535ng 3:3 m.» <5 mo mva 2: ma SE .3 «.53... NO 151. Q. ..0 mmml mono? «mod . mxa mN Qm ..Q Md v.0 mum m6 LED Q. .9 8.6:? 44a .6 my... xa 97 Table VII. Acidity Functions of PAA at Various Ionic Strengths at 25°C. Ionic Strength x 102 9K1 9K2 9K3 9.95 2.26 i 0.04 4.90 i 0.02 8.04 i 0.02 6.954 2.22 i 0.04 4.93 i 0.02 8.07 i 0.02 5.144 1.97 i 0.04 4.91 i 0.02 8.11 i 0.02 5.104 2.18 i 0.04 4.94 i 0.02 8.14 i 0.02 3.224 2.01 i 0.04 4.96 i 0.02 8.23 i 0.02 2.492 1.91 i 0.04 4.95 i 0.02 8.22 i 0.02 1.963 2.13 i 0.04 5.02 i 0.02 8.32 i 0.02 1.876 1.93 i 0.05 4.99 i 0.02 8.32 i 0.02 0.00 2 2.0 5.11 i 0.04 8.69 i 0.05 98 .gpmeaeum 68:68 mo.o 88 <: 8x c. .u: ...M low. “A“ Low. n v.— c. {\j IrocN 101 momp ..86 AAA ..68 xoom .FA=-;Aeouz .PN-P a .Aepmweaeu PAUAA»_8=< to 3668888: .AmAAazA com.“ NA.m- m.m~ ‘ mo.o- ocean: 46 mg Am~.e ~4.8- 8.FN no.9- 86 open» 102 acid and H3P04 (126). The protonation reactions are entropy stabilized. This is expectable since we are replacing two species of opposite charges by one species with a smaller overall charge. 3.2. COMPLEXATION STUDIES OF PAA 3.2.1. SPECTROSCOPIC STUDIES 3.2.1.1. Visible Spectra - Aqueous solutions of manganese chloride are yellow-green. Upon the addition of PAA, the solution turns pink. A possible explanation for this color change is given by Cotton and Wilkinson (127). The majority of manganese complexes are high spin, in octahedral field this configuration gives spin forbidden as well as parity forbidden transitions, thus accounting for the extremely pale color of such compounds. In tetrahedral environments, the transitions are still spin forbidden, but no longer parity forbidden, the energy of the transitions is about 10 times stronger, and the compounds have a noticeable pale yellow-green color. These results show that there is complexation between the metal ion and PAA. Interaction of PA with copper ion was indicated by changes in the absorbances of copper ion solutions upon addition of PA. The molar absorptivity at Amax of 780 1 5 nm is 10. When PAA is present at a mole ratio PA/Cu of 1.8, in basic medium (pH higher than 9), Ama is at 765 z 3 nm with a molar absorptivity x of 30. 103 3.2.1.2. Sodium-23 NMR - When enough PA was added to a 2 x 10'2 M_solution of NaCl at pH above 9, to make the PA/Na+ ion mole ratio equal to 8, the sodium-23 resonance shifted slightly downfield from 0.36 to -0.05 ppm. At the same time, the line-width increased from 11.2 to 21.2 Hz. While those changes are small, they indicate unambiguously the formation of a weak sodium complex. 3.2.1.3. Phosphorus-31 NMR - Results similar to the sodium-23 NMR ones were observed. The phosphorus-31 resonance shifts slightly downfield from -l3.3 to -l4.9 ppm for a sodiumzPA mole ratio change from 0 to 10. The change in chemical shift of a 0.3 M_PAA - 0.3 M_MgClz solution v§_pH is shown in Figure 20 and compared to the results previously obtained for PAA alone. The plot presents several interesting points. At pH above 9, when the potentiometric results indicate quantitative complexation of magnesium ions, the chemical shift is -l4.7 ppm compared to ~13.3 ppm for the free PA. We have seen that protonation of the acetic acid moiety results in an upfield shift of the phosphorus-31 NMR signal, while protonation of the phosphoric acid moieties causes a downfield shift. At a pH of 9, PAA is completely deprotonated and, therefore, becomes a tridentate ligand. Formation of the Mg2+ complex affects all three sites and the result is a downfield phosphorus-31 shift. Secondly, at pH between 3 and 6, the NMR signal of a PAA solution moves downfield when MgBr2 is added, indicating evidence 2- of complexation with the LH ion (i.e., formation of a mono- protonated complex). 104 .18 w.» 8.5.38 558: E A. .6 2; s: a ma .8 :23 829.... 5.38885 .8 2.5: In gawkmncnm. dd- < 8mm 6 2A -8. a 3.9.8.. .. «no.8 _ Nanozvoo ...o «Sm 110 Table IX. Half-wave Potential of the Cd(II)/Cd Couple in the Presence of PA v§_Tha11ium Amalgam. [50315/11-3 log [PA] E1/2 (mV) 0.00 194.2 5.022 -2.299 145.3 10.06 -1.997 133.3 15.09 -1.821 124.8 20.13 -1.595 118.5 25.16 -1.599 113.3 30.20 -1.520 109.0 35.23 -1.453 105.3 40.27 -1.395 100.5 45.30 -1.344 98.0 50.34 -1.298 94.3 lll -FL-B F2 (L) - 1([21 1 = 82 (45) assuming we only have a lzl and a Zzl complex. A plot of F0 (L) v§_[L] will be a steeply rising curve, however, as [L] approaches zero, the graph will have a limiting slope of B], and an intercept on the Fo (L) axis of l. A preliminary value of B1 is obtained. A plot of F] (L) y§_[L], on the other hand, will have a limiting slope as [L] tends to zero, of 82 and an intercept on the F1 (L) axis of B]. A confirmative estimation of B] is possible and, in addition, a preliminary value of 82 is obtained. The values of B1 and 82 found are given in Table X. The precision in the determination of B] and hence of 62 is limited by the previously mentioned assumption that we have a high ligand to metal mole ratio. 3.2.3. POTENTIOMETRIC STUDIES OF COMPLEXATION 3.2.3.l. Sodium Ion Electrode Study - Due to the low stability of the sodium-PA complex and the high charge of the PA ions, the experiment was performed at l.0 ionic strength. It has been shown (lBl) that at high ionic strength, the activity coefficient is much less sensitive to the variations of the ionic strength. Thus, in our experimental conditions, the change in ionic strength was small enough during the course of the titration to be neglected in a first approximation. The value of the formation constant is 3.3 (Table X). 3.2.3.2. Glass Electrode Study - Figure 23 represents a coulometric titration of PAA and PAA with MgBrz. In the latter 112 o.P ll o—c ¢.o ll H U mo.o wuo.o II H P.o a m._ “Nazvcx mop F.o a m.m mm.o Aszvcx mo_ one unz mo.o a om.N oo.o a mm.~ mo.o a ~m.~ mo.o a om.~ mo.o a m~.o Azozvcx no. 35 a $5 85 ... 3...” No.0 .. 85 No.0 a o: 85 a m: :5: 9: om mam (mag no: maz .mcoH papa: cur: mmxmpaeou <L=u :omumguw» owguosopzou- .mm mgammu mFZwJ<>50w m N _ O _ _ _ \...,.i.. . _ .IgnvAvmvl >Em_uw< .13!) .6 I l I .. .. .e. loo—I >Ehhum< ., - . ... _ . Io 6.49%on . e . 3,5 . . . . m . . . <3 . . . mum... _ loo. JOON < 5.8:? 39):; 00. .mm ac=m_a 118 constants of the mono- and deprotonated complexes are 1.0 x 103 and 4.0 x 105, respectively. By analogy with the cadmium-PAA system, similar complexes were assumed in the magnesium-PAA case. However, our potentiometric results disagree with this hypothesis and show that the magnesium- PAA complexes are different from the cadmium-PAA ones. These results show evidence of completely deprotonated and monoprotonated magnesium-PAA complexes. These complexes have one or no charge. In these conditions the change in ionic strength during the titration is negligible, and the value of the ionic strength can be equated to the concen- tration of supporting electrolyte. A species distribution plot at an ionic strength of 0.05 is shown in Figure 26. It can be seen that for a 1:1 mole ratio of PAzmagnesium, approximately 50% of the magnesium (or PAA) is complexed at pH = 7, which is the physiological pH. The formation constants of the PAA-calcium ion complexes were determined at 25°C at an ionic strength of 0.05, and compared to the values of the magnesium ion complexes. A similar study was performed with strontium and barium ions. The stability of the complexes decreases with increasing radius of the metal (see Table X). The complexation of sodium ion has been studied by the same method (Table X). At 0.078 ionic strength, the sodium-PA complex has a formation constant of 27. Owing to the low ionic strength, this value is compatible with the formation constant of 3.3 found by sodium ion electrode measurement at 1.0 ionic strength. 119 .gumcaepm u_coH mo.o be Emumsm uz-<1 :3 o ).l_ . .Cl. 6 (" . t). 7 :tva 7 &::3 LT;3.'I $: )1 :y s:.;: 659'. N'- Tgl L {IN I|.; Y 'L‘av’ :t'll 3'\ i3 ” 0. v0 JH :osfiu 9LT-VYI - v {FNST' «909:? 3110: I I 9’ $5693.12 UL) gua:tur 3: OULQ'L " 2L ll.) vii-Q9 ' OOOHOJJ)OYO D ( 1 I V 73 3! :P". 3‘. ‘ Z T‘. or N St I I1 I 1’ u L‘: 3 ANAI‘QL F: :1 so ... ... ... ... ... o. a.YN//1.... 25:7...(o: .55 ..J J J. UTNOIIIT'IC l. ..I (( Unb.IoIOOJTHII‘eJfl.B°!.T-I§: aJJoToIéaf. azuaTaltq. W. J 1.9 U U J .J J {...-ONES oI(:.(.. N::2:NKKNNN...NNIO....UTU :Usqa... TN0.0T..45:.0 (0. NT 1..-C...) se.:-E§1-Nlt.|8£‘ . N OI .432...VKEVNN:.VN.\N\N\%NN.. abT LCOONJOTG 2.0.1.1 (...-5.121.111.94! Ir... 1: II! a In 5:21 . OH. O TIEIvTCIIO 01....11TNI1058122TNQ. 7 0.. I152HI0R£I¢ I. .- .a.‘. 8‘. ..Q. .3 9.... 3d 01 o1‘..:..13... ’TT1'33 Jv J..Jo {JV t ta4J1Jfi. 2 JIXI?‘T(.T(1.-3,.3.ITHV 3.. A :TTI TTTnIa‘uoITTY TTTT. T T 1..-1......IJTJV Jo.o.JTJo.J.oJTJTJ ...NIIVQIIITTGQ. VVJIITITNTNCY‘TTLthTIY L L L 9.. "(To-\(TI "150:9... S‘s-e... It.‘s.|s...l I ... V II ..LILSTNTNTNITVTNTN? NT NT. . L? b' LOTF vHNE‘OLOTYLCIblCFLmC LIE..E?|.L.IFL.1EEFP I .. ...: ..llr‘ OFCIINFLBIVIT‘Fépabtf SAX-.753 L ROLF ..EECI.C...UICCI.U.LCLn.Ir 2." ...(u. .v (LIL flux-.29... ....CnhwjbrrnsCIPCCCFVDKOCXCOKDKCINFCSQPIBP. “FEDTSquTh39‘NCo-Tog N C (15109 ECPLEK. .....r. RC; ....» (EFL... 017.1. I. o 1 7.25 2 .2751 111 122 2.. 1.L.w.u9|u1’. 3 222 b.- b l- .0 0 o o 0 6 .... o 2 1 e. . ”- o 1 2 ‘. O. 2 ‘ c a 1 2 3 CE 140 Subroutine EQN 8.2.2.2. 1111111111 5.0088000 .... 000000.000 NNN 0 .00 0 ....It ... 2.. T T v- 0 o a .I 1... o R. ... 3. PVT 1 on. 0 22577.01: T .9 ()I I l o 7.0.1,??501 . .-.... UCD .... I. 1.! 18356100512 1’. ..1T. . I. (... 0000000000 J o v 0‘ o 9 w .0 “120.05.00.50. 9.5. ’V' m 2. o V. 6 I. van. 0 ‘1 IU.U It" 6 c . ‘ (.5 " ‘.b‘ II. \ o 0 (v... N u L (.2 u fiT ~25 a. an. 0 0 V 3.362200 A :0. 3.... r...“ 101T... ......o... . 1 T ... f. 11 0.3%. or} To... LA. I. /K 000000000 . x. o ...... II WC.- . a 1. ssgo-O‘Sz’ .... L o - Colo 0 al.- V .l. .5): 1..."..NIZ... \L. ... .l .1 ”.....pug.’ -..“ “...? I. fvhfl‘) .. '7 O .(V’ 05%. 0., 0 . n17 fLo .. n. YELHLFV... .. ‘.227I.21..1.. 9.5 1a.... "ERGO... n.- ...-...... C... 30...! J HF 1C1COT .5 tit—...: Ni O bru OI r o n T. s .- Tr. 1...]! o. “(131% O ‘14.}? 3.1.9 «I v Iii! J L H . . ... ... 23.301.73.505 . . .7 ) v . v v T511... . .... l 10%.-536:9... u. . . 0.l!) . .0 N ... h )1 ... 2S..7‘.°I.9.3‘. ..IC 1....5. .. OTOMIWT.M.2.I 9 00.00.0000 I u h . .... ... ..v. a O 0 5.5.0:. 71.11.... v v. 0TH... N.T FTFTOT D -—- ... ..........1............ DD. 1.11),: .l D‘s-To .J 1 . 0. h.o.b.b.t.$.§b“f. 1': 1-..... 2 159.7 f r‘: .1 . L coco-0000000 5T .. h 0 OK 1 Cl T..l1rT-L‘O I. 3 ... ... .’o 1.”! ‘110 I LLNPN N 0.. .. I no ...; P D. T v 9 l. . ... 1 ...........IT!.TT< .. ... 0 n V 45' T...) o 1 T.... TIT! I. I . .. 1 a. ..JD '“31’ 0 do.“ £5.09:U' , , . L a. c .. o X 0" .J .u ...‘F.‘N1N H OT 1.. on 1 ... a. ... H. ... IV 31:54! . ...!L.L°t.n\<..l|. 7 O T. L a. V I»! I. S V 3 c. 0 01.2.7. 0 NH .... .. a... ... o ... I. I. . .15 ... .v ....C‘... .9.‘...§7 .. .J .-.. INV . 0.. ..4 . L‘ . .1 I1 ... u I 1 b 0557...... .7133? . . .J 3.1!. .V s .. A .NIJAOTZ D? a... T .. .J 1.).. .J a. 10 .J A N 21.39 1160.....‘a. ....J .vU ....v 6 LTHOCOC1H. FIG. I. ’2 T T; T I ..T 7 .. ...... none... 1 . 0 J 9.: 3 o 9 .1... .. .....1‘. .o o -2... s T o I ...... ... .l O T D.C.... .- 5b.. ('5... ...H . ... ...) v ...T Tn.-~((.. 7.. .T .30 ...1 00 J . .... n )0 0 v ... I u I o .. ...l. a O ..I. 81.5151... 0 is .6. 1.? I: not 1;... ‘9. 1x... rt 2‘. (r A U 0 s . . ...: (L). I... . . ...... h F" Vito! . 00 .59.. 20 ‘2 I To. I 0 . D l 1. v .I 0 Pl...‘ 1. v o P 15.10;»... 6“. 0|. T 0 TI ‘2 0.! 9. ..TI .1 .u . ..I .‘v I I. . c. . .120 n. .9. .K on..." 13 . .. .... f. 0 .... o D... 1 . L oVC l/LCSD... o I. I...(.v-.va.oo3t.o oIvTv.I .U C. ...). ‘0 an 7 .o . .... ...... (...... r... 3 L 0 T1 (....TC... .- Ia ...lT 7.1/0 «1.1.9 0P1 511 o... 0.1. I ... IL .0 o a ... o. . . . . . . . . . . .0 .0 . .1 ll. . ... I ...! .... 01:81.. 19(‘39I 107.510 . 0 T o .I o 1'- 0 V I .l. 0.. ... u 2. FI.E..0.C..€I.:. ....I...9......V. a... at... v. ...v ..VC.N .1. ... lt..’.../h..9 1.09.1... -0 z 0. I! o s 0.7...." N ‘1 ... 00.00000. N. . 3.7.. CN’Clo.(1 0 .vv ... .- .II...N3’II:IT1 .0 3.2.1.6. .(ITCI..¢. 1... I33. 2,. .v 2". D. on. o o . 0 Ca 5511...: . . . .4.-. ...... D a). .. .r. c... .... 5.. ... ... ...1..V ... cl .1... .1)! J..(.ISO o . 159.1103TT TT .....J).T '0! ...I.‘ o . v. .110... ... ... ... ...“ ... ... ... ... ... 95:. L. 7 SIN/[:1 1......:CJ.....8.. U U Uqu.)T’(l:.. T I. C. /.ool(c.Ctfi.1 .....1.)... n.U.T .15... ..UQ...To/OT.. U U U UT U U U U U 3 ... o -.N ..«JNH...)T...(1 N: ......\NNNP.n.1-J..1J1.J ..J..U:Obv.noc..1T..3....... ...-... ....I..£..NIQ.II...S.....NOIL1L.U.....V‘V‘VNNV.&VNKNNNNNNVN o c... ....T -...OCN. ...: ... 6...... (3:23.19 1.. 1X. ... (.... .. .. I... b. H. 0 (31C... 3T1TNI1§.2TN3§H..‘P 11.101.- 1011110....1‘10131‘10. . .11 ... .1N........D ... ....r 711.. 3 Jv 1T JT .. :JJlJ? ... .... 1.. ...(1. .3321... ....TZTTTTTI.‘ . TTTT TTII. .TTI..(I..I9.’.JTJY Jo JT..JTJT Jo-JTITJ 2 ... ...!!VVNNHIIDTia VIJIIV NTQTNTYLaLILTCTL L LCI.((T(.\.......L.. S 5!... .....NS‘ .. 32.1.. .. ..LILTTTVTNTNY ‘TTNY ‘TITNTNT... 1.5.3333336 1...... UOTGV 0.. ...Y. N5. .UCOUTT ..OLLECLH!..I2! .7? L11....... .. .. .. :E'Cfit.hu.0|..v.l$‘.rlr.0.t.r .PFUQRKFBLI1TIEOIOBEOF1015tthEL o . o . . . . . . . . . . . «Pt-1.! (tn. 9?. 1.9 TGCNNIJFSCFECT((UPCiTLVOYDQZLfiIE CFFTCGTQTN‘. INT.TN’iNE’EXgI'ECPESFECVCFCFCS.... B P (.L111§11l L... I151. ..Z'. N C 1 7 I. ... 1 ..U. 10. H H... ..u n P. uh “... .... B 5 .1... 0. I. nu ..(a ..L....J... .. ... a 1...... a 1 ’ I. 1‘2 0. ... o. 2 ..., o. 2 ... O «up...un!.w...«..nt’.u.......u7 . 1 ... ... . o ......c...’.\..... SE Q 00000000000 61‘.- 00‘ 111.1. APPENDIX C APPLICATION OF THE COMPUTER PROGRAM MINIQUAD 76A TO THE DETERMINATION OF EQUILIBRIUM CONSTANTS FROM POTENTIOMETRIC DATA I41 I42 APPLICATION OF THE COMPUTER PROGRAM MINIQUAD 76A TO THE DETERMINATION OF EQUILIBRIUM CONSTANTS FROM POTENTIOMETRIC DATA C.I. PROGRAM MINIQUAD 76A This program (136) determines the formation constants from potentiometric data. Up to 20 equations involving 5 reactants can be considered at the same time, and up to 3 electrodes can be used to determine the concentration of free species. The equation constants are calcuiated for the reaction + aA + bB + CC +’ AaBch (65) with the formation constant A B C f = [aa bbC] c (66) [A] [B] [C] In such a case the reaction LHZ’ + Mg2+ Z MgLH (67) with the fbrmation constant M LH Kf(MgLH) = "9 LH (68) becomes L3” + H+ +Mg2+ Z MgLH (69) with . M LH or , K M LH Kf (MgLH) = f( g ) (71) K3IPAA5 l43 With the computer program MINIQUAD 76A, the formation constants can be either held constant or refined in the calculations. The coulometric (volumetric) titration data consist of the measured electrode potential as a function of the time during which the titrant was generated (the volume of titrant added). In these acid base titrations, one of the mass balance equations describes the concentration of species involving dissociable hydrogen ions. The hydroxide ion is considered as a "negative proton", and the proton concentration in a basic titrant is equal to the negative value of the hydroxide ion concentration. C.2. DATA INPUT INSTRUCTIONS FOR MINIQUAD 76A C.2.l. 1 card /20A4/ : descriptive title C.2.2. 1 card /815/ : LARS, NK, N, MAXIT, IPRIN, NMBEO. NCO, ICOM. LARS is an indicator for the data points to be considered in the refinement: with LARS=l all the data points are used, with LARS=2 alternate points, with LARS=3 every third point, 239. (last points on all titration curves are always used). NK is the total number of formation constants. N is the number of formation constants to be refined. MAXIT is the maximum number of iteration cycles to be performed: with MAXIT=0 and according to the values of JPRIN and JP (see below) the residuals on mass balance equations and/or the species distribution are evaluated for the given formation constants and conditions. T44 IPRIN=O is normal; IPRIN=l monitors the progress of the refinement at each cycle; IPRIN=2 produces an additional listing of the experimental data at each titration point. NMBEO is the total number of reactants (mass balance equations) in the system under consideration. NCO is the maximum number of unknown concentrations of free reactants; if NCO=O the whole job is abandoned before refinement. ICOM=O is normal; with ICOM=l data points are eliminated before the refinement if the corresponding block of the normal equation matrix is found to be not positive-definite. C.2.3. 1 card /3Fl0.6, 8X,IZ/ : TEMP, ADDTEMP, ALPHA, NOTAPE TEMP is the reaction temperature in °C. ADDTEMP is the titrant temperature in °C. ALPHA is the coefficient of cubical expansion for the solvent used, °C’]. NOTAPE=O is normal; NOTAPE=l reads values for EZERO and SLOPE (see below) from device TAPE3. This allows calibration curve data to be calculated and used in the same computer run. C.2.4. NK cards /FlO.6,715/ : BETA(I), JPOT(I), JQRO(J,I) (NMBEO values), KEY(I) The formation constants are expressed in exponential notation B] = BETA(I) - ioJP0T(I) - JQRO(J,I) (J=l, NMBEO) are the NMBEO stoichiometric coefficients of the 1th species with formation constant Bi' The order of coefficients is arbitrary, except that those referring to reactants l45 of which the free concentration is determined potentiometrically must come last. Such a choice implies that a progressive integer number (from 1 to NMBEO) is assigned to each reactant. KEY(I) is the refinement key of the 1th formation constant: with KEY=O the formation constant is not refined and with KEY=l the formation constantis refined. C.2.5. The following set of cards for each titration curve: 1 card /1215/ : NMBE, JNMB(I) (NMBE values), NC, JP(I) NMBE is the number of reactants (mass balance equations) involved in the titration curve. JNMB holds the integer numbers previously assigned to the NMBE reactants involved. NC is the number of unknown free concentrations at each point of the titration curve; the number of concentrations experi- mentally determined (j;g;, the number of electrodes) is NEMF=NMBE-NC. JP contains integer numbers corresponding to selected reactants: in the subroutine STATS the formation percentages relative to these reactants will be calculated, depending on the value of JPRIN. 1 card /5AlO/ : REACT(I) (NMBE values) REACT contains the names of the reactants, listed in the same order as JNMB. 146 1 card /4IS/ : JEL(I) (NEMF values), JCOUL JEL holds the number of electrons transferred at each electrode. If the decimal cologarithm of concentration (QLQL- pH) is to be read in, put JEL(I)=O. JCOUL=0 is normal; JCOUL=l if the total volume of the solution does not change during the titration (gEQL, coulometric experiments). l (or 2) card(s) /8FlO.6/ : TOTC(I) (NMBE values), EZERO(I) (NEMF values), ADDC(I) (NMBE values), VINIT TOTC contains the initial number of millimoles of reactants in solution; the order of reactants is the same as in JNMB. EZERO(I) holds the standard potential of the 1th electrode (mV); the value is ignored if JEL(I)=O. ADDC contains the molar concentrations of titrant solutions (there is one for each reactant); the order of the reactants is the same as in JNMB. VINIT is the initial volume of the solutions (cm3), and should correspond to the volume expected at the temperature of the TITRANT. 1 card /8FlO.6/ : SLOPE (NEMF values). SLOPE contains the slopes of the calibration curves for the species measured, in units of mV per decade of concentration, the value is ignored if JEL(I)=O. l47 cards /15,8F8.3/ one for each point of the titration curve: LUIGI, TITRE(I) (NMBE values), EMF(I) (NEMF values) LUIGI=O is normal, LUIGI=l indicates the end of a titration cruve, LUIGIl, the amount and type of tables and/or graphs is determined by the values contained in JP for each titration curve. C.2.7. 1 card /15/ : NSET NSET=l for another set of formation constants - items C.2.l. - C.2.4., C.2.6. and C.2.7. only -; NSET=O for another complete set of data, NSET=-l for the termination of the run. I48 DATA SAMPLE C.3. {til . b . =4 .... (Lot-.....coz {a TA - ,1. 0......Is..b ... '- \4 rd .- I... ooo;!t1..).la$.~ 1 ...» K n. ,I L L $In . . .9 »\.L O ..r. a. S. I- n. .- ..1 I... .1 ‘4... 3...... 0 ..II I. . v .0 n O c n t I 0 O ' ,- Or f 1 v Ls 1 J .o . . . 1 a ‘ a O .. Co 0. 35.06 6. 1 ‘ ‘ 0 Q . 'IQ ‘- u r. .c, c .9 O ... 55.‘t fl.Sluggaala,93l.’°’55$550~951.5553‘.!I.9959956‘9I.° ...U.J b.UaJ.Lf,-. ..U-LPI. .1? .0325. .306. . 4°}! co; urtdsolqsflucat oooocoonocoou...-cocoounocucocoa-......ooou o‘c‘921“‘oz‘io*.‘:;§7\*i-i‘31. VVVVVVV'VJ"'VUV'V'VV'VV"'"l'ivvvvvvvv‘vv E04615... .t‘.... E. ..’...1.a... 3:53,! .5 ...; :3....7.5v....17n . 1339!... 3355.... OOOOOOOOOOOOOOOCOIDIOIOOOOO00.00.000.000... .1 filth; . Clio-t. ‘5‘... Lt.r~‘....b..tt 7695...?» 4.75.. e: 31“... . .. 2 {Fitzbl’ ......B ctiuz. 51.5.5351... 27....h5A . 112.335.33.79... ~L2~s§¢1..101..¢5Qo..110c511¢a.2110 . o o o o o o o o o o . ..7.as.3.a7:1 .2. o? aSJaiacr P... v7? ’5. .51’5’552‘ 315613.“.‘3 . .ru)? 05.. . s. ....5 ~. .... cat's-1192.: . u, :2... II): ~. . L. 28:6. - Stole-Er. IO0..........OI......‘II.......‘OCOOOCCO... uOCC—LOOJDCD... ruck-s... C50...CSC:.O...C¢..USO¢. 650.25.553.3— C .3. 3.102113... 05.51.... ’3)..), .4323, . o3..77-J;37 1,.."J)..,’ .d):”.d 137.1531... 5.7.9; .9 b n. b O71 77 3.31. 00 Q: .c..1....)~2223333 b 1113111102....111: 00.00.000.000000000000000... 0 5L ..Ltlfikslkkt 5....ch5 C‘s SI. LLL,5L.L.~cr-.0 3.3 Q ......IOOCOIIOIOOOIOOIQII......‘OI........I 7 ...-... ......I p. a. .. oo o. .- —‘: o. .... o. ..J... ..vnc-b.o.OQ-..Jqoo ..5.‘-.Jo varufJQOoo QC 0 I. . sf. . :5 5"» ......» ...; etc LL.L.§L LLU‘ hLLI- ..va tt‘tt‘t‘OIL; BIBLIOGRAPHY IO. 11. 12. 13. 14. 15. 16. 149 BIBLIOGRAPHY (a) A. I. Popov, Pure and Applied Chem. 31, 3 (1975). (b) w. J. Dewitte, R. C. Schoening and A. I. Popov, Inorg. Nucl. Chem. Lett., 12, 251 (1976). (c) J. 5. Shih and A. I. Popov, Inorg. Nucl. Chem. Letters, 1_3_, 105, (1977). F. Alder and F. C. Yu, Phys. Rev., B2, 105, (1951). M. Ellenberger and M. Villemin, C. R. Acad. Sci. Paris, Serie B, 266, 1430 (1968). P. D. Dougan, S. N. Sharma, and D. Llewelyn Williams, Can. J. Phys., 51, 1047 (1969). E. M. Dickson and E. F. u. Seymour, J. Phys. Chem., g, 666. (1970). FL G.Bryant, J. Magn. Reson., 6, 159 (1972). J. A. Magnusson and A. A. Bothner—By in "Magnetic Resonance in Biological Research," A. Franconi, Ed. Gordon and Breach, London, 1969, p. 135. L. Simeral and G. E. Maciel, J. Phys. Chem. §Q, 552, (1976). J. H. Swinehart and H. Taube, J. Chem. Phys. 31, 1579, (1962). N. A. Matwiyoff and H. Taube, J. Am. Chem. Soc., 99, 2796, (1968). R. D. Green and N. Sheppard, J. Chem. Soc. Faraday Trans. II §§, 821, (1972). F. Toma, M. Villemin, and J. M. Thiery, J. Phys. Chem., 11, 1294, (1973). A. D. Covington and A. K. Covington, J. Chem. Soc. Faraday Trans. 1, 11, B31, (1975). R. N. Butler, J. Davies, M. C. Symons, Trans. Faraday Soc., 66, 2426, (1970). B. N. Chernyshov, V. V. Kon'shin, Koord. Khim. g, 365, (1976). E. C. Ashby and R. C. Arnott, J. Organometal. Chem., 13, l, (1968). 17. 18. 19. 20. 21. 22. 23. 24. 25. 26. 27. 28. 29. 150 (a) Nouveau traite de chimie minerale, Pascal, Masson et Cie Editeurs, 1958, tome IV, p. 170, and references therein. (b) Handbuch der anorganische chemie, Gmelin, Berlin, Verlag Chemie GMBH, 1939, 27, teil B, p. 140 and references therein. (c) Ikeuchi and Krohn, Acta Chem. Scand. 23, 2232, (1969). (d) D. Bryce-Smith, Inorg. Synthesis, Vol. 6, p. 9. (e) H. J. Abendroth, A. Peters, w. Fischer, Rev. Chim. Min., u, 556, (1974). (f) L. V. Gorbunov, V. N. Deciatnik, S. R. Raspopin, K. I. Trifonov, Zh. Neorg. Khim., 12, 3093, (1974). (g) u. D. Treadwell and Th. Zurrer, Helv. Chim. Acta, 15, 1271, (1932). H. F. Halliwell and S. C. Nyberg, Trans. Faraday Soc., 59, 1126, (1963). F. L. Cook, T. C. Caruso, M. P. Byrne, C. N. Bowers, D. H. Speck, and C. L. Liotta, Tetrahedron Lett., 46, 4029, (1974). M. A. Neuman, E. C. Steiner, F. P. van Remoortere, and F. P. Boer, Inorg. Chem., 14, 734, (1975). R.D. McLachlan and V. B. Carter, Spectrochim. Acta, Part A, 39, 1171, (1974). A. Pullman, C. Giessner-Prettre, and Y. V. Kruglyak, Chem. Phys. Lett., 35, 156, (1975). F. A. L. Anet, J. Krane, J. Dale, K. Daasvatn, and P. 0. Kristiansen, Acta Chem. Scand., 21, 3395, (1973). B. K. Leong, T. 0. Ts'o and M. Chenoweth, Toxicol. Appl. Pharmacol., 22, 342, (1974). P. B. Chock, Proc. Nat. Acad. Sci. US, 62, 1939, (1972). B. K. Leong, Chem. Eng. News, 4, 5, (1975). B. Dietrich, J. M. Lehn and J. P. Sauvage, and J. Blanzat, Tetrahedron Lett., 2885 and 2889, (1969). B.Dietrich, J. M. Lehn and J. P. Sauvage, and J. Blanzat, Tetrahedron, gg_1629, (1973). J. M. Lehn and J. P. Sauvage, J. Amer. Chem. Soc., 22, 6700, (1975). 30. 31. 32. 33. 35. 36. 37. 38. 39. 40. 41. 42. 43. 44. 45. 46. 47. 48. 151 w. E. Morf and w. Simon, Helv. Chim. Acta, 54, 2683 (1973). J. M. Lehn, Structure and Bonding, l5, 2 (1973). J. J. Christensen, 0. J. Eatough, and R. M. Izatt, Chem. Rev., 24, 351 (1974). J. 0. Lamb, R. M. Izatt, J. J. Christensen, and D. J. Eatough, "Thermodynamics and Kinetics of Cation - Macrocycle Interaction," Chapter in Chemistry of Macrocyclic Compounds, G. A. Melson, ed., Plenum, N.Y. in press. 0. N. Bhattacharyya, C. L. Lee, J. Smid, and M. Szwarc, J. Phys. Chem., 55, 608 (1965). J. S. Shih, Ph.D. Thesis 1978, Michigan State University. J. G. Hoogerheide, Ph.D. Thesis 1978, Michigan State University. G. J. Templeman and A. L. Van Geet, J. Am. Chem. Soc., 54, 5578 (1972). D. H. Live, 5. I. Chan, Analytical Chemistry, 42, 791 (1970). M. S. Greenber , R. L. Bodner and A. I. Popov, J. Phys. Chem., .22, 2449 (1973). U. Mayer and V. Gutmann, Structure and Bonding, 12, 113 (1972). (a) V. Gutmann and E. Hychera, Inorg. Nucl. Chem. Lett., 2, 257 (1966). (b) V. Gutmann, "Coordination Chemistry in Nonaqueous Solvents" Springer-Verlag, Vienna (1968). M. S. Greenberg, D. M. Nied and A. I. Popov, Spectrochim. Acta, 29.6.. 1927 (1973). Y. Cahen, Ph.D. Thesis 1975, Michigan State University, East Lansing, Michigan. A. I. Popov, Pure and Applied Chem., 42, 3 (1975). A. I. Popov, "Polyhalogen Complex Ions" in Halogen Chemistry, vol. 1 V. Gutmann, ed., Academic Press, London and N.Y. (1967), p. 253. R. H. Erlich and A. I. Popov, J. Amer. Chem. Soc., 93, 5620 (1971). M. Greenberg, Ph.D. Thesis 1975, Michigan State University, East Lansing, Michigan Y. M. Cahen, P. R. Handy, E. T. Roach, and A. I. Popov, J. Phys. Chem., 22, 80, (1975). 49. 50. 51. 52. 53. 54. 55. 56. 57. 58. 59. 60. 61. 62. 63. 64. 65. 66. 67. 68. 69. 152 T. C. Farrar and E. 0. Becker, Pulse and Fourier Transform NMR, Academic Press, N.Y. and London, (1971), p. 65. R. H.)Er1ich and A. I. Popov, J. Amer. Chem. Soc., 93,5620 1971 . J. S. Shih and A. I. Popov, Inorg. Nucl. Chem. Letters, 19, 105 (1977). E. Kauffmann, personal communication. K. Balasubrahmanyam and G. J. Janz, J. Am. Chem. Soc., 92, 4189 (1970). P. Nylen, Ber., 979, 1032 (1924). M. I. Kabachnik and T. A. Mastryukova, Izvest. Akad. Nauk., S.S.S.R., 163 (1953). A. A. Shvets, 0. A. Osipov, and A. M. Shakirova, J. Gen. Chem., USSR, 92, 2588 (1967). E. M. Popov, M. I. Kabachnik, and L. S. Mayants, Russian Chem. Rev., 99, 362 (1961). L. J. Bellamy and L. Beecher, J. Chem. Soc., 475 (1952). L. J. Bellamy and L. Beecher, J. Chem. Soc., 729 (1953). . J. Harvey and M. G. Horning, J. Chromatogr., 29, 65 (1973). D R. J. Maile, Jr. G. J. Fischesser and M. M. Anderson, J. Chromatogr. , 132,366 (1977). H 1.1. . Sohr, J. Electroanal. Chem. Interfacial E1ectrochem., . 188 (1966). D. Balzer and H. Lange, Colloid Polym. Sci., 259, 643 (1975). D. Balzer and H. Lange, Colloid Polym. Sci., 255, 140 (1977). w. Althoff, M. Fild and H. P. Rieck, Z. Naturforsch. B. Anorg. Chem. Org. Chem., 919, 153 (1976). G. A. Gray, J. Am. Chem. Soc., 99, 2132 (1971). J. G. Riess, J. R. Van Nazer, and J. H. Letcher, J. Phys. Chem. 11, 1925 (1957). T. Bottin-Strzalko and J. Seyden—Penne, J. Chem. Soc.,Chem. Commun., 22, 905 (1976). L. L. Burger, J. Phys. Chem., 62, 590, (1958). 70. 71. 72. 73. 74. 75. 76. 77. 78. 79. 81. 82. 83. 84. 85. 86. 87. 88. 89. 90. 91. A. A. Elesin, A. A. Zaitsev, S. S. Kazakova, and G. N. Yakovlev, 153 Radiokhimiya, 14, 541 (1972). A. A. Elesin, A. A. Zaitsev, S. S. Kazakova, and G. N. Yakovlev, Nuc1. Sci. Abstr., 25, 18030 (1973). S. M. Lambert and J. I. Hatters, J. Am.Chem. Soc., 29, 4262 (1957). .1. Soc., 7_9, 3651, (1957). I. Hatters, S. M. Lambert and E. D. Loughran, J. Am. Chem. J. Mao, E. Robishaw, Biochemistry, 14, 5475 (1975). J. F. Cannelongo, U.S. Patent #3,370,029. Th. Auel, Ger. Offen. #2,505,435. Anonymous, Neth. Appl. #6,508,180. G. T. Bottger and A. P. Yerington, U.S. Dept. Agr., Bur. Entomol. Plant Quarantine E-863, (1953). E. E. Nl—Ca—i 84 F. manor-TI Rogers, A. A. Patchett, Ger. Offen. #2,318,254. Rogers, A. A. Patchett, Ger. Offen. #2,318,254. . Clark,U.S. Patent #4,016,264. . Herrin, J. S. Fairgrieve, U.S. Patent #4,052,439. . Schleicher, w. R. Roderick, Ger. Offen. #2,208,787. Overby, R. G. Duff, J. C. Mao, Ann. N.Y. Acad. Sci., 310 (1977). J. A. Boezi, in press. J. Hay, S. M. Brown, A. T. Jamieson, F. J. Rixon, H. Moss, 0. A. Dargan and J. H. Subak-Sharpe, J. Antimicrob. Chemother., 9, Suppl. A, 63 (1977). S. S. Leinbach, J. M. Reno, L. F. Lee, A. F. Isbell, and J. A. Boezi, Biochemistry, 15, 426 (1976). B. H. Fox, J. Antimicrob. Chemother. 9, Suppl. A, 23, (1977). 1. H. Maugh II, Science. L92. 128. R. J. Klein and A. E. Friedman-Kien, Antimicrob. Agents Chemother., _7_, 289 (1975). H. G. Adams, E. A. Benson, E. R. Alexander, L. A. Vontver, M. A. Remington, K. K. Holmes, J. Infect. Dis. 199 (Suppl.): A151-A159 (1976). 92. 93. 94. 95. 96. 97. 98. 99. 100. 101. 102. 103. 104. 105. 106. 107. 108. 109. 110. 111. 154 E. L. Goodman, J. P. Luby, M. T. Johnson, Antimicrob. Agents Chemother. 9, 693 (1975). A. E. Friedman-Kien, A. A. Fondak, and R. J. Klein, J. Invest. Dermatol., 55, 99 (1976). D. D. Gerstein, C. R. Dawson, and J. 0. 0h, Antimicrob. Agents Chemother., 2, 285 (1975). R. F. Meyer, E. D. Varnell, and H. E. Kaufman, Antimicrob. Agents Chemother., 9, 308 (1976). J. F. Fitzwilliam and J. F. Griffith, J. Infect. Dis., 199, Suppl. A221, (1976). N. L. Shipkowitz, R. R. Bower, R. N. Appell, C. W. Nordeen, L. R. Overby, W. R. Roderick, J. B. Schleicher, and A. M. von Esch, Appl. Microbiology, 25, 264 (1973). R. J. Klein and A. E. Friedman-Kien, Antimicrob. Agents Chemother., 2, 289 (1975). J. Roboz, R. Suzuki and G. Bekesi, and R. Hunt, Biomed. Mass. Spectrom., 4, 291 (1977). B. A. Bopp, Ch. 8. Estep, D. J. Anderson, Fed. Proc., 95, 939 (1977). . R. Herrin, J. S. Fairgrieve, R. R. Bower, N. L. Shipkowitz, . Mao, J. Med. Chem., 29, 660 (1977). . F. Lee, K. Nazerian, S. S. Leinbach, J. M. Reno, J.A. Boezi, . Natl. Cancer Inst., 55, 823 (1976). . Nylen, Z. Anorg. Allg. Chem., 295, 33 (1937). . Warren and M. R. Williams, J. Chem. Soc., (B), 618 (1971). . Denham, U.S. Patent #2,629,731. . R. Naqvi and P. J. Wheatley, and E. Foresti-Serantoni, . Chem. Soc. A, 12, 2751 (1971). . M. Reno, L. F. Lee, J. A. Boezi, Antimicrob. Agents Chemother., 3. 188 (1978). A. Unni, L. E1135, and H. Schiff, J. Phys. Chem., 52, 1216 (1963). T J L P S P. Nylen, Ber., 529, 1036 (1924). H R J J 1 D. Duv111ier, Ann. Chim. Phys. 29, 327 (1881). B. Schmidt, Ann., 252, 265 (1892). 112. 113. 114. 115. 116. 117. 118. 119. 120. 121. 122. 123. 124. 125. 126. 127. 128. 129. 130. 131. 155 Handbook of Chemistry and Physics, Chemical Rubber Company, ed., (1967-8) p. F-4. M. M. Crutchfield, C. F. Callis, R. R. Irani and G. C. Roth, Inorg. Chem., 1, 813 (1963).‘ R. Jones and A. Katritzky, J. Inorg. Nucl. Chem., 15, 193 (1960). J. H. Letcher and J. R. Van Wazer, J. Chem. Phys., 44, 815 (1966). J. H. Letcher and J. R. Van Wazer, J. Chem. Phys., 45, 2916 (1966). J. H. Letcher and J. R. Van Wazer, J. Chem. Phys., 45, 2926 (1966). J. H. Letcher and J. R. Van Wazer, Chapters 2 and 3 in Phosphorus 31 Nuclear Magnetic Resonance, M. Grayson, E. J. Griffith, ed., Interscience Publishers, N.Y. and London, (1967). K. Moedritzer, Inorg. Chem., 5, 936 (1967). Carbon-13 NMR for Organic Chemists, Levy Nelson, ed., Interscience (1972), p. 23. J. C. Tebby, Organophosphorus Chemistry, vol. 1, Specialist Periodical Reports, The Chemical Society, ed., London (1970), p. 294. G. Mavel and M. J. Green, Chem. Comm., 742 (1968). J. C. Tebby, Organophosphorus Chemistry, vol. 2, Specialist Periodical Reports, The Chemical Society, ed., London (1970) p. 252. R. G. Bates, Determination of pH Theory and Practice, J. Wiley and Sons, Inc., N.Y., London, Sydney, (1964), p. 53. H. S. Harned and B. 8. Owen, The Physical Chemistry of Electrolytic Solutions, Reinhold Publishing Corp., N.Y., 3rd edition, (1958), p. 165. L. Meites, Handbook of Analytical Chemistry, McGraw-Hill Book Co., N.Y., San Francisco, Toronto, London, Sydney, Tst edition, (1963), p. 1-21. C. Cotton and J. Wilkinson, Advanced Inorganic Chemistry, Interscience, 3rd edition, p. 848. J. J. Lingane, Chem. Rev. 29, 1 (1941). D. D. DeFord and D. N. Hume, J. Am. Chem. Soc., 29, 5321 (1951). S. Fronaeous, Acta Chem. Scand.,4, 72 (1950). H. S. Harned and B. B. Owen, the Physical Chemistry of Electrolytic Solutions, Reinhold Publishing Corp., N.Y., 3rd edition, (1958), p. 738. 156 132. L. Meites, Handbook of Analytical Chemistry, McGraw-Hill Book Co., N.Y., San Francisco, Toronto, London, Sydney, lst edition, (1963), p. 1-46. 133. P.-H.)Heube1 and J. Hoogerheide, "Constant Current Coulometer," 1977 . 134. J. L. Dye and V. A. Nicely, J. Chem. Educ., 49, 443 (1968). 135. F. A. L. Anet, J. Krane, J. Dale, K. Daasvatn, and P. 0. Kristiansen, Acta Chem. Scand., 22, 3395, (1973). 136. A. Sabatini, A. Vacca, and P. Gans, Talanta, 21, 53 (1974). 137. D. D. Traficante, J. A. Simms and M. Mulvay, J. Magn. Reson., E, 484 (1974). 138. D. A. Wright and M. T. Rogers, Rev. Sci. Instr., 44, 1189 (1973).