SPECTROSCOPIC STUD’uES OF SONIC SOLVATION EN NO‘NAQUEOUS MEDIA Dissertation for the Degree of Ph. D. MiCHEGAN STATE UNIVERSETY MARK S. GREENBERG 1974 is”! . L I13 2? A ,3 Michigan 5;- . 13% University a m . . . \ . pm ‘ a '. .' Y? r I :_ w . I, I - ‘ ' I! . _ , | . t . g 5 - “ h This“! to certify that the - thesis entitled " 'l SPECTROSCOPIC STUDIES OF IONIC SOLVATION IN NONAQUEOUS MEDIA ' O 3:; Mark az‘éfien . “ ' ’1 . _ ,i A .. .V - {h - . presempd ‘ ‘5 . . "h t \ ( > I“ ' l a; has been accepted » ar-‘f 3 a of the requilém , k. ' egre” 7 i I: . 79" / '. ‘7 professor e. r" 4-.- q .‘ ' _ ‘ I t . ‘— V 33' 1' u" 0 ‘ - . ' LI .’ . . . 5 . - . ‘3 fl 1‘ r a I. I I ABSTRACT SPECTROSCOPIC STUDIES OF IONIC SOLVATION IN NONAQUEOUS MEDIA By Mark S. Greenberg Sodium-23 NMR measurements have been performed on several sodium salts in eighteen nonaqueous solvents as a function of salt concentration. The chemical shifts for solutions of sodium tetraphenylborate and perchlorate exhibited little or no concentration dependence, whereas chemical shifts of corresponding solutions of sodium thiocyanate, bromide and iodide showed marked concentration dependence. The static nature of the 23Na shifts in the former case is proposed to be indicative of either free Na+ ion or solvent separated ion pairs. The 23Na shifts in the latter case are proposed to be indicative of contact ion pair formation. A plot of the infinite dilution 23Na chemical shifts in these solvents against the Gutmann donor numbers of these solvents yielded a straight line. Hence, the magnitude and direction of the 23Na chemical shift reflects the relative donicity of these solvents. A quantitative evaluation of ion pair formation constants suggests that the formation of contact ion pairs is strongly influenced not only by solvent dielectric constant but also by the "donicity" (solvating ability) of the solvent. The interpretations obtained @\ Mark S. Grecnberg from the 23Na chemical shift correlate well with the data obtained from electrical conductance measurements. Competitive solvation studies for the Na+ ion in binary solvent mixtures of nitromethane, acetonitrile, hexamethylphosphoramide, dimethylsulfoxide, pyridine and tetramethylurea were monitored by 23Na NMR. Generally, these studies reflected the relative donicity of each solvent in a given solvent pair where the solvent of higher donicity was preferentially contained in the inner solvation shell of the Na+ ion. However, in binary solvent mixtures of pyridine with dimethylsulfoxide and tetramethylurea, the donicity of pyridine seemed to be repressed. Infrared studies of the former mixture suggest that strong solvent-solvent interactions disrupt the associative structure of DMSO molecules resulting in enhanced donicity for DMSO in these mixtures. Because tetramethylurea exhibits structure similar to that of DMSO, the same solvent—solvent interactions are proposed to enhance its donicity. The slight upfield shift of the 23Na resonance with increasing concentration for solutions of sodium perchlorate was reexamined by 23 5 Na and 3 Cl NMR, Raman and infrared spectroscopy. Linear upfield shifts of the 23Na resonance were noted for solutions of NaClO4 in methanol, ethanol, water and formic acid whereas non-linear upfield shifts were observed in the other solvents studied. In pyridine, acetonitrile and tetrahydrofuran, a new Raman band at N 470 cm-1 in addition to the other C10 - bands at 456, 626, and 935 cm"1 was 4 observed. With increasing concentration, the 470 cm-1 band increased Mark S. Greenberg in intensity relative to the 456 cm-1. Hence, these bands were proposed to be indicative of bound and free C104“, respectively. The line- width of the 35Cl resonance was identical within experimental error 35 (30 :_5 Hz) for all solutions of NaClO as was the Cl chemical 4 shift (-lO40 1.5 ppm). These data suggest that contact ion pair formation of NaClO4 does occur in the solvents studied; moreover, the interaction is weak. In the hydroxylic solvents, the energy of interaction is proposed to be less than kT, since the 23Na Shift varies linearly with concentration. SPECTROSCOPIC STUDIES OF IONIC SOLVATION IN NONAQUEOUS MEDIA By Mark Sf Greenberg A THESIS Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Chemistry 1974 ACKNOWLEDGEMENTS The author wishes to thank Professor Alexander I. Popov for his guidance, encouragement, friendship and counseling throughout this study. He also wishes to thank Professor Joseph A. Caruso of the University of Cincinnati for his unique introduction to research, his friendship and many good times. Gratitude is also extended to the Department of Chemistry, Michigan State University, the National Science Foundation and the National Institutes of Health of the Department of Health, Education, and Welfare for financial aid. Appreciation is extended to Dr. Richard Bodner for numerous enlightening discussions and experimental assistance during the early part of this work. The present members of the "group", Paul Gertenbach, Duke DeWitte, Yves Cahen, Robert Baum, DavidDeBrosse and John Thompson, should be acknowledged for their gifts of friendship and enthusiasm, and an incessant willingness to lunch at McDonalds. Special thanks to Mr. Patrick Kelly for interfacing the Raman instrument and providing program PEAKSBF and to Wayne DeWitte and Robert Baum for their expenditure of time in proofreading this thesis. Professor Stanley R. Grouch, Crouchl, Crunch, Couchy, etc. is acknowledged not only for his helpful suggestions as second reader, but for his outstanding sense of humor. The author would also like to thank Professor Robert Hammer and the students of Honors Chemistry, 1970-1971, for providing a most unusual and rewarding teaching experience and some important friendships. Special thanks are given to Messrs. Eric T. Roach, A. Wayne Burkhardt and Frank Bennis, without whose cooperation the NMR investigations would have been much more difficult. Deep appreciation is extended to my wife, Roni, for her love, patience and encouragement throughout the years of graduate study. To her and to our families, I dedicate this thesis. ii TABLE OF CONTENTS Chapter I. HISTORICAL NUCLEAR MAGNETIC RESONANCE . INFRARED SPECTROSCOPY RAMAN SPECTROSCOPY . CONCLUSIONS II. EXPERIMENTAL PART SALTS SOLVENTS . SAMPLE PREPARATION . INSTRUMENTAL MEASUREMENTS Nuclear Magnetic Resonance Infrared Spectra. Laser Raman Spectra . Conductance Measurements Data Handling . . . . . . III. A SODIUM-23 NUCLEAR MAGNETIC RESONANCE AND ELECTRICAL CONDUCTANCE STUDY OF CONTACT ION PAIRS IN NONAQUEOUS SOLVENTS INTRODUCTION . RESULTS AND DISCUSSION . . . IV. STUDIES OF PREFERENTIAL SOLVATION OF THE SODIUM ION IN MIXED SOLVENTS BY SODIUM-23 NMR INTRODUCTION . . . RESULTS AND DISCUSSION . CONCLUSIONS . . . . . . . . . iii Page . 12 . 15 . l6 . 16 17 . 20 . 20 22 22 23 . 23 . 24 . 25 51 O 52 78 Table of Contents (Continued) v. A SPECTROSCOPIC STUDY OF CONCENTRATED SOLUTIONS OF SODIUM PERCHLORATE - THE NATURE OF THE UPFIELD CHEMICAL SHIFT INTRODUCTION . RESULTS AND DISCUSSION . CONCLUSIONS v1. APPENDICES I. DESCRIPTION OF COMPUTER PROGRAM KINFIT AND SUBROUTINE EQN FOR THE CALCULATION OF ION PAIR FORMATION CONSTANTS BY THE NMR TECHNIQUE 11. DESCRIPTION OF COMPUTER PROGRAM KINFIT AND SUBROUTINE EQN FOR THE RESOLUTION OF OVERLAPPING RAMAN AND INFRARED BANDS III. DESCRIPTION OF COMPUTER PROGRAM PEAKSBF FOR THE RESOLUTION OF OVERLAPPING RAMAN AND INFRARED BANDS . IV. DESCRIPTION OF COMPUTER PROGRAM SHEDLOV FOR EVALUATION OF CONDUCTANCE DATA . VII. LITERATURE CITED iv 80 81 . 107 . 108 . 112 . 115 . 123 . 126 LIST OF TABLES Table Page 1. Salvation Band Frequencies of Alkali Metal Ions in Nonaqueous Media . . . . . . . . . . . . . . . . . . 9 2. Key Solvent Properties . . . . . . . . . . . . . . . . 18 23 3. Na Chemical Shifts vs. 3.0 M_Aqueous NaCl . . . . . . 27 4. Ian Pair Formation Constants for Various Sodium Salts by 23Na NMR . . . . . . . . . . . . . . . . . . . 50 5. Variation of the Sodium-23 Resonance as a Function of Solvent Composition for Binary Solvent Mixtures . . . . 53 6. Isosolvation Points for Salvation of the Sodium‘ Ian in Binary Solvent Mixtures. . . . . . . . . . . . . 63 7. Variation in the Frequency of the Sodium Ian Salvation Band for 0.50 M NaBPh4 Solutions in DMSO- Pyridine Solvent Mixtures . . . . . . . . . . . . . . . . . . . 66 8. Variation of the S- O Stretching Frequency of DMSO for DMSO- Pyridine Solvent Mixtures . . . . . . . . . . 69 9. Covington Treatment of Preferential Salvation . . . . . 77 10. 23Na Chemical Shifts for Sodium Perchlorate Solutions in Nonaqueous Media . . . . . . . . . . . . . . . . . . 82 11. Association Constants of Sodium Perchlorate from Conductance Studies . . . . . . . . . . . . . . . . . . 88 12. Computer Analysis of Sodium Perchlorate Solutions in Tetrahydrofuran and Acetonitrile by PEAKSBF and KINFIT O O O O O O O O O O O O O O O O O O O O O I O O 96 13. KINFIT Analysis of Raman Spectra of Sodium Perchlorate Solutions on Acetonitrile from 900- 950 cm‘1 . . . . . . 100 List of Tables (Continued) 14. 35Cl Chemical Shifts for Sodium Perchlorate Solutions in Nonaqueous Media . 15. 35Cl Linewidth for Sodium Perchlorate Solutions in Nonaqueous Media . vi . 103 . 104 Figure 10. 11. 12. LIST OF FIGURES Sodium-23 Chemical Shifts of Various Sodium Salts in l,l,3,3-Tetramethylurea and N,N-Dimethylformamide . Sodium-23 Chemical Shifts of Various Sodium Salts in Sulfolane and Acetone . . . . . . . . . . . . Sodium-23 Chemical Shifts of Various Sodium Salts in Pyridine and l,l,3,3-Tetramethy1guanidine Sodium-23 Chemical Shifts of Various Sodium Salts in Tetrahydrofuran. . . . . . . . . . . . . . . Conductance Curve far Sodium Iodide in 1,1,3,3-Tetramethy1guanidine. . . . . . . . . . . . Conductance Curve for Sodium Iodide in Pyridine . Sodium-23 Chemical Shifts of Various Sodium Salts in Methanol and Ethanol . . . . . . . . . . . . . . . Sodium-23 Chemical Shifts of Various Sodium Salts in Propylene Carbonate and Dimethylsulfoxide . . . . . . . Plot of Infinite Dilution Sodium-23 Chemical Shifts. versus the Donor Number of the Solvent. . . . .-. . Sodium-23 Chemical Shifts of Sodium Iodide in Sulfolane, Formamide, Acetone and Dimethylsulfoxide . . . Variation of the Chemical Shift of the Sodium-23 Resonance as a Function of Solvent Composition for Binary Solvent Mixtures of Nitramethane with Acetonitrile, Tetramethylurea, Dimethylsulfoxide, Pyridine and Hexamethylphospharamide. . . . . . . . Variation of the Chemical Shift of the Sodium-23 Resonance as a Function of Solvent Composition for Binary Solvent Mixtures of Acetonitrile with Tetramethylurea, Dimethylsulfoxide, Pyridine and Hexamethylphospharamide . . . . . . . . . . . . . . . . . vii Page . 26 . 33 . 34 . 35 . 37 . 38 . 39 . 41 . 43 . 58 59 List of Figures (Continued) 13. 14. 15. 16. 17. 18. 19. 20. 21. 22. 23. Variation of the Chemical Shift of the Sodium-23 Resonance as a Function of Solvent Composition for Binary Solvent Mixtures of Hexamethylphospharamide with Tetramethylurea, Dimethylsulfoxide and Pyridine . . . . . . . . . . . . . . . . . . . . . . . . 60 Variation of the Chemical Shift of the Sodium-23 Resonance as a Function of Solvent Composition for Binary Solvent Mixtures of Tetramethylurea with Dimethylsulfoxide and Pyridine . . . . . . . . . . 61 Variation of the Frequency of the Sodium Ian Salvation Band for 0.50 M_NaBPh4 Solutions in Dimethylsulfoxide-Pyridine Solvent Mixtures . . . . . . 67 Variation Of the 5-0 Stretching Frequency for Dimethylsulfaxide—Pyridine Solvent Mixtures . . . . . . 7O Covington Plat for the Binary Solvent System Dimethylsulfaxide-Pyridine . . . . . . . . . . . . . . . 76 Sodium-23 Chemical Shifts of Sodium Perchlorate- Solutians in Nonaqueous Media . . . . . . . . . . . . . 86 Vibrations of the C104- group as a Function of Symmetry . . . . . . . . . . . . . . . . . . . . . . 91 Raman Spectra of 1.0 M_Sodium Perchlorate Solutions in Tetrahydrofuran and Acetonitrile fI‘Om 400-500 C111 0 o o O a o a o o o o o o o o o o o o 93 KINFIT Analysis of 1.0.M_Sadium Perchlorate Solution in Tetrahydrofuran from 400-498 cm‘l. . . . . . 94 PEAKSBF Analysis of 1.0 M Sodium Perchlorate Solution in Tetrahydrofuran from 430—498 cm'l. . . . . . 95 Raman Spectra of Sodium Perchlorate Solutions in Acetonitrile from 900-950 cm-l. . . . . . . . . . . .101 viii LIST OF NOMENCLATURE, ABBREVIATIONS AND SYMBOLS Contact Ion Pairs. Pairs of ions, linked electrostatically, but with no covalent bonding between them. Solvent Shared Ion Pairs. Pairs of ions, linked electrostatically by a single, oriented solvent molecule. Solvent Separated Ian Pairs. Pairs of ions, linked electrostatically but separated by more than one solvent molecule. ACN: Acetonitrile PC: Propylene Carbonate THF: Tetrahydrofuran DMF: Dimethylfarmamide TMU: 1,1,3,3-Tetramethylurea DMSO: Dimethylsulfoxide HMPA: Hexamethylphospharamide TMG: l,1,3,3-Tetramethy1guanidine A = equivalent conductance of the solution = equivalent conductance at infinite dilution S = the Onsager slope = OAO + B a = the relaxation effect = (8.205 X 10-5)/(DT)3/2 B = the electrophoretic effect = (82.43)/n(DT)1/2 D = solvent dielectric constant ix List of Nomenclature (Continued) ['11 ll L. ll m ll absolute temperature viscosity concentration Ele + E ale + 02 degree of dissociation 2 association constant mean activity coefficient CHAPTER I HISTORICAL NUCLEAR MAGNETIC RESONANCE One of the central problems in bath aqueous and nonaqueous solutions of electrolytes is the characterization of the interaCtions that exist between the constituent ions and among the ions and the solvent. Chemical literature contains more than nine hundred electrical conductance studies of electrolyte solutions since classically this was one of the first approaches and still continues to be one of the more important methods for characterization of ionic equilibria in solutions. Other techniques which have been employed to study complexatian and salvation in electrolyte solutions include spectrophotometry in the visible and ultra-violet region, distribution equilibria, ultrasonic relaxation techniques, potentio- metry, vibrational spectroscopy and magnetic resonance (1-6). Nuclear magnetic resonance (NMR) has become a powerful tool for the investigation of electrolyte solutions. There have been many studies of the chemical shift and relaxation times of protons of the solvent molecules or in the solvated species (e.g., tetraalkylammonium salts) (7-15). However, relatively wa studies of the magnetic resonances of nuclei other than protons present in the ions themselves have been reported, although the chemical shifts and line widths of the nuclear resonances would yield information about both ion-ion and ion—solvent interactions. All the alkali metal and halide ions possess at least one isotope with a magnetic nucleus, i.e., 7Li, 23Na, 39K, 87Rb, 133Cs, 19F, SSCI, 79'813r and 127I. Deverell and Richards (16) 2 studied the chemical shifts of 23Na, 39K, 87Rb and 133Cs nuclei in aqueous solutions of alkali halides and nitrates as a function of salt concentration. The concentration dependence of the chemical shifts were attributed to interactions between the cations and anions in solution. These same workers surveyed the chemical shifts of 35Cl, 81Br and 1271 nuclei in aqueous solutions of alkali halides and determined that direct cation-anion collisions were the predominant cause of the chemical shifts of the halogen resonances in solutions of potassium, rubidium and cesium halides. In lithium and sodium halide solutions the cation-water-halide interaction was of prime importance (17). WennerstOrm, gt, a}: (18) measured NMR linewidths of 35C1, 793i, Slat and 127 I nuclei in aqueous solutions of various substituted ammonium, phosphonium.and sulfonium salts. The observed line broadening was attribUted to anion—solvent interactions due to the structure making or breaking action of the cation. Noting that 35Cl chemical shifts reflect the solvent environment of the Cl' ion, Langfard and Stengle (19) studied competitive salvation of Cl' ion in acetanitrile-water and dimethyl- 35C1 and 7Li line- sulfoxide-water mixtures. Bryant (20) monitored widths far concentrated LiCl solutions (1-15 M) and suggested that changes in the linewidth were indicative of ion-pairing phenomena. Maciel, gt, 9;, (21) observed 7Li chemical shifts of dilute solutions of lithium bromide and perchlorate in water and eleven organic solvents. Recent studies by Akitt and Downs and work in this laboratory reveal the frequency of the 7Li resonance to be quite sensitive to the environment (22-24). Halliday, 333 213 (25) investigated 133Cs chemical shifts of cesium halides, hydroxide and nitrate in water and seVeral nonaqueous solvent. The shifts were found to vary non-linearly with salt concentration. The degree of nan-linearity depends upon the dielectric constant of the solvent. Hence, the shifts reflect cation- anian collisions. Carrington, gt, 31. (26) observed linear shifts in 133Cs resonance as halide salts are added to a solution of CsCl and postulated contact ion pair formation to be responsible for these shifts. Bloor and Kidd (27) determined the 39K chemical shifts in a number of aqueous electrolyte solutions as a function of concentration and observed both upfield and dawnfield chemical shifts due to random cation-anion collisions. The sodium-23 nucleus is also well suited for such.nmr studies, and therefore, it was the subject of investigation in this thesis. The relative sensitivity of 0.1 at constant field with respect to the proton indicates that measurements can be obtained with fairly dilute solutions. The small natural linewidth of the sodium ion resonance (~15 Hz) permits the use of high resolution NMR eqUipment. Finally, 24 esu cm2 renders this nucleus the large quadrupole moment of 0.1 K 10' a sensitive probe of the neighboring electronic environment. The first detailed 23Na nmr investigation was performed by Jardetzky and Wertz (28) who reported an Observed broadening of the 23Na resonance line as a fhnction of concentration and anion which they attributed to an interaction of the 23Na nuclear quadrupole with an 4 electric field gradient caused by the formation of weak complexes of sodium with the added anion. Rechnitz and Zamochnick (29) evaluated the formation constant of sodium ion with several acid anions. They plotted these formation constants vs. the Jardetzky-Wertz linewidths and observed a linear relationship. Eisenstadt and Friedman (30) determined relaxation times for aqueous solutions of sodium perchlorate and chloride finding that the former has a marked effect on the 23Ma relaxation time whereas the latter did not. In a study similar to that Of Jardetzky and Wertz, Griffiths and Socrates (31) found the linewidth of the 23Na resonance for several inorganic salts to be prOportional to the concentration, indicative of ion pair formation. However, they likewise did not observe any variation in the 23Na chemical shifts in their systems. Richards and Yorke (32) studied 79Br, 81Br and 23Na resonances for aqueous and nonaqueous solutions of sodium, calcium and cesium bromide. The results indicate that for sodium and calcium bromide solutions relaxation arises from ion-solvent interactions whereas ion-ion interactions are important for cesium bromide solutions. In 1968, Bloor and Kidd (33) reported 23Na chemical shifts of sodium iodide solutions in fourteen nonaqueous solvents as a fUnction of concentration. The concentration dependence of the observed chemical shifts was attributed to cation-anion interactions; moreover, the observed range of the chemical shifts was related to changes in the paramagnetic term of the general screening equation. A fair correlation was found between the magnitude of the chemical shift extrapolated to infinite dilution in a given solvent and the Lewis basicity of the solvent. In order to extend these measurements to new solvents and also to determine the influence of different anions on 2:I’Na chemical shifts, Erlich, 333 21, (34) studied 23Na chemical shifts of sodium tetraphenyl- borate, perchlorate, thiocyanate and iodide in several nitrogen and oxygen donor nonaqueous solvents. In the case of the first two salts, the chemical shifts were found to be independent of concentration, while for the last two, varying degrees of concentration dependence were noted. In this latter case, it was assumed that the chemical shifts were influenced by the formation of contact ion pairs. A linear 23Na chemical shift correlation was found between the magnitude of the in a given solvent and the Gutmann donor number* of that solvent. Templeman and Van Geet (35) measured 23Na chemical shifts and relaxation times for aqueous Solutions of sodium tetraphenylborate, perchlorate, hydroxide and chloride and postulated that short-lived collisional ion pairs are responsible for the observed chemical shifts. By monitoring the 23Na chemical shift as a function of added water, Van Geet (36) determined the hydration number of sodium ion to be between three and four. Canters (37) examined the interactions of glyme ethers with the sodium ion as a function of temperature, anion, solvent and viscosity. *Gutmann's donor numbers (96) are the enthalpy of complex formation between the given solvent and antimony pentachloride in 1,2-dichloraethane solution S + SbClS w S'SbCls Gutmann used the term "donicity" when referring to the donor ability of a solvent. 6 Since the Na+ ion is of biological significance, several studies were pursued to further elucidate the magnitude of interaction of the sodium ion with biologically important molecules. The interaction of Na+ with biological tissues has been reported by Cope (38) and Ratunno (39). The camplexing of ionophores which function as mobile carriers of alkali cations in membranes, such as valinomycin, manensin, 23Na NMR (40). A good nigericin and monactin, were studied with correlation was observed between the 23Na resonance position of the 1:1 complex and the stability constant of the complex. Sodium-23 NMR relaxation time measurements have been used to investigate the formation of weak complexes of sodium ion with some aminopalycarboxylic acids (41), soluble RNA (42), sodium-potassium transport adenosinetriphophatase (43), some phosphate-containing compounds of biological interest (44) and interactions with cysteine, aspartic acid and citric acid in aqueous solution (45). The magnitude of the change of relaxation time yields a qualitative indication of the strength of the sodium ion interactions with the ligand. Andrasko and Forsen (46) applied pulsed fourier transform 23Na NMR to the study of the binding of sodium ion with simple carbohydrates. Cerasa and Dye (47) employed 23Na NMR to study the exchange rate of the sodium ion between two environments in solution; ethylenediamine and a hexaoxadiamine macrobicyclic camplexing agent (cryptate). Below the coalescence temperature of 50°C, two well defined resonance absorptions were Observed, indicative of free and complexed sodium ion. Shchari, gt: a}, (48) monitored complexation of sodium ion by dibenzo- l8-crown-6 in dimethylfarmamide but were unable to observe two peaks 7 because the linewidth of the complexed Na+ ion was very broad, although the line shape analysis indicated that exchange was slow. INFRARED SPECTROSCOPY While studying the far infrared spectra of tetrabutyl- and tetrapentylammonium chlorides and bromides in benzene solution, Evans and Lo (49) noted bands at 120 cm-1 and 80 cm.1 respectively, which could not be attributed to a vibrational mode of the solute or the solvent. Since the band position was dependent both on the mass of the cation and anion, it was ascribed to the direct cation-anion ion pair vibration and constituted the first report Of an ionic vibration in solution. At the same time, Edgell, 33: 31: (50) observed far infrared bands of alkali metal tetracarbonylcobaltates and pentacarbonylmanganates in tetrahydrofuran solutions. These bands likewise did not arise from either the solute or the solvent. After extending these studies to such solvents as dimethylsulfoxide, pyridine and piperidine, it was concluded that the band position was a function of cation and solvent. Hence, the observed band represented the alkali ion vibrating in a cage (51). These vibrations were named "salvation bands". POpov and ca-warkers extended these far infrared studies to a wide range of nonaqueous media (52—60) and observed these salvation bands to exhibit strong cation dependence and much weaker solvent dependence. The bands, which follow Beer's Law within experimental accuracy, are intense, but broad, with band widths equal to or greater 1 than 50 cm' . Isotopic substitution far the cations or solvent indicated that band frequencies vary inversely with the change in mass Of the cation or the solvent. Generally, no anion dependence is noted in solvents of high polarity or donicity; however, in some solvents of low polarity or donicity, such as tetrahydrofuran (50,51), acetone (56) and propylene carbonate (60) some anion dependence is noted. Such dependence is postulated to be indicative of contact ion pairing in solution. The frequencies Of the alkali metal ion salvation bands in several solvents are presented in Table 1. Generally, the Li+ salvation band occurs at N400 cm-l, Na+ and NH4+ at mZOO cm-l, K+ at mlSO cm-l, Rb+ at m120 cm-1 and Cs+ at mllO cm-l. These data are indicative of the strong cation dependence Of the frequency Of these bands. A unique electrolyte, sodium tetrabutylaluminate, which is soluble in solvents of very low polarity, has been the subject of extensive investigation. Tsatsas and Risen (61) reported two concentration 1 for this salt in 1 dependent far infrared bands at 195 cm-1 and 160 cm- cyclahexane and tetrahydrofuran solutions in addition to a 202 cm- Raman band in the farmer which is attributed to sodium ionic motion. In addition to a 1H nmr study (8), Olander and Day investigated salvation of sodium by tetrahydrofuran in the system THF-NaAlBut -cyclohexane 4 by monitoring the v stretching frequency of THF in the region COC 900-1150 cn’1 (62,63). Edgell's group (64) investigated the effects of different solvents and cations on the infrared spectrum of metal tetracarbonylcobaltate salt (MCo(C0)4) solutions. The 1890 cm“1 band (C-O stretch) was faund to be quite symmetrical in dimethylfarmamide, dimethylsulfoxide, and wet tetrahydrofuran, indicative of a symmetrical "solvent-surrounded" mm Rm cm em mm.em am mm mm mm.mm.mm Hm Hm mm.Hm Hc.oc Oucoaowom wvH OOH ova mvH NmH MNH omH OHH mNH mmH omH +mU +Dm +M + OON NON oHN CNN NNN vHN emz AHIEOV moflocoswouu mcmm coaum>fiom owa mad VON NON VNN CNN OON owH me ooN omH m2 ONv omm va mHv mom oov va ONv va va now “A ocfioflusa vwo< proo< ocopoo< ocowflaouhxthachw>tH Ocowfiaonpxthtfixcpoth Ocowwfiohhxth ocfixomasaasesefia ObfixomH5meOHmHo ooaxomfisoaseSoeAa ocwkuxm oceceeoafla cc-ocflxomaanAEOoeaa cmaawoavxsmuuoe muco>fiom «five: msoosvmcoz :fi mcoH Hope: flamxa< mo mofiocoscoam wcmm :ONpm>Hom .H CHQMH 10 00 mm mm mm mm mm Nam mad va va CON owH owa oov mmm mmm Nom vwm 0mm apmcopnmu ecoHNQOHm ocfiwfihxmonofisutN ocflvfiexmaxzuoaflotv.m ocflwflhxmaxsuoEthV.N ocfivfihxmaxnuozum onwwfiaxmfixnuoztv AEODCMucouv H OHDMH 11 ion environment. However, in pyridine, piperidine, dry tetrahydrofuran and dimethoxyethane the band is split into two components above and below the 1890 cm—1 band. These data indicate increasing asymmetry about the anion resulting from contact ion pairing. A temperature study Of NaCO(C0)4 in tetrahydrofuran solutions enabled the complex band envelope around 1890 cm-1 to be resolved into four components indicative Of two anion environments assigned as contact and solvent separated ion pairs (65). Recently Barbetta and Edgell (66) examined the infrared spectrum of thallium tetracarbonylcobaltate in seven solvents as a function of temperature. Only a single ion site was found in dimethylfarmamide, dichloromethane and dimethylsulfoxide solutions. Several kinds of sites were found in tetrahydrofuran, acetonitrile and nitromethane solutions including free ions, solvent separated ion pairs, contact ion pairs and triple ions. Tsatsas and Risen (67) observed far infrared bands for lithium, sodium and calcium ions in ethylene-methacrylate ionic capolymers. These same workers also investigated far infrared bands of alkali metals in cyclic polyether "crown" compounds in pyridine and dimethyl- sulfoxide solutions (68). The results presented in this thesis are applications of the aforementioned studies. A more extensive historical discussion of salvation studies by infrared spectroscopy can be found in the Ph.D. theses of B. W. Maxey (69), J. L. Wuepper (70), P. R. Handy (71) and M. K. Wong (72). 12 RAMAN SPECTROSCOPY Recently, with the advent of laser excitation Raman spectroscopy, the methods Of vibrational spectroscopy to probe ion-ion and ion- salvent interactions are very much in the forefront in the studies of electrolyte solutions. For axyanions such as perchlorate, nitrate and sulfate, the splittings and intensities of the anion Raman lines have been suggested as probes Of the disposition of solvent molecules around the ions and indicators of the presence of contact or solvent separated ion pairs. Such studies were originally pursued in aqueous solutions of metal oxyanion salts. Hester, 33, 31, (73) investigated aqueous solutions of indium sulfate, nitrate and perchlorate and found in all solutions a broad, polarized Raman band at 400 cm.1 which probably resulted from hydrated In3+. In both the indium sulfate and nitrate solutions, three new polarized Raman lines appeared which were accounted for in terms of C3v and C2V symmetry for bound sulfate and nitrate anion respectively. NO new bands due to bound perchlorate were observed. Hester and Plane (74) examined Cu2+, Zn2+, Hg2+, Mg2+, In3+ and Ga3+ sulfate, nitrate and perchlorate solutions and noted a new polarized line at 360-400 cm-1 in each solution. Because, for a given cation, this band was independent of anion, it was suggested that the origins of these lines are in some symmetric farms of vibrations within the hydration sheaths of the solvated metal ions. These same workers studied saturated aqueous metal ion-oxyanian solutions and concluded that cation-nitrate complexes are much more common than the corresponding sulfate complexes, while there is no evidence for the existence of perchlorate complexes (75). 13 Several research groups have suggested that the magnitude of splitting of the v2(E') band of D3h free nitrate into v4(Bl) and v1(A1) lines characteristic of bound C2v nitrate may be used as a measure of the strength of the cation-nitrate interaction (76-78). Using this criterion, Hester and Plane (75) proposed the following . . . . 4+ order of cation-nltrate Interaction strength: Th > In3+ > Cu2+ > Hg2+ > Ce3+ > Ca2+ > Zn2+, A13+, Ag+ > Na+, K + , NH4+ (75). While studying the Raman spectra of the nitrate ion in aqueous solutions of cadmium and zinc nitrate, Irish, gt: a}: (79) proposed spectroscopic criteria for solvent separated and contact ion pairs. A splitting of the v3(E') band in the 1300-1500 cm.1 region and the absence of splitting in the 720 cm.1 region is proposed to be indicative 3 A contact ion pair interaction features an additional loss of of an interaction of the type M+(H20)nN0 (solvent separated ion pair). degeneracy of the v4(E') mode resulting in two bands at 720 and 740 cm-1. After examining the 1050 cm-1 region of aqueous solutions of calcium, cadmium, zinc, and silver nitrate in detail, Janz, gt: a}, (80) further proposed that a nitrate band at 1036 cm"1 is indicative of a contact ion pair whereas a band at 1041 cm'1 is indicative of a solvent separated ion pair. Hence, these detailed Raman studies of aqueous solutions of metal nitrate salts resulted in several criteria to serve as guidelines for the identification of ionic species in solution. In order to evaluate these criteria in nonaqueous media, detailed Raman studies of metal nitrates in acetonitrile were pursued, principally by Janz. In addition, solvent bands could also be monitored to gain further insight as to the disposition of solvent in these solutions. 14 Balasubrahmanyam and Janz (81) examined silver nitrate solutions in acetonitrile and noted the data to be in accord with spectroscopic criteria for contact ion pairing. In the N03” stretching frequency region, bands at 1036 and 1041 cm—1 are observed, indicative of free and bound nitrate respectively. The relative concentration dependence of these bands permits calculation of a, the fraction of free nitrate ions, as a function of concentration. The band contours of the CEN and C-C stretching frequencies of the solvent are each resolved into two components, indicative of salvation of the metal ion. This study confirms the conclusions of an earlier infrared investigation of complexes of AgNOS-CHSCN and AgNOS-ZCHSCN by the same workers (82). Addison, et: 31, (83) examined infrared and Raman spectra of solutions of zinc, cadmium and mercury(II) nitrates in acetonitrile and found the nitrate spectra to be consistent with a strong perturbation of nitrate ions by solvated cations. Only in the case of mercury(II) was a metal-oxygen vibration frequency observed. Vibrational studies of electrolytes in liquid ammonia have recently received much attention. Gardiner, 33, 313 (84) examined lithium and ammonium nitrates in liquid ammonia and noted large changes in the N-H stretching region, 3000-3500 cm-l, which were correlated with the metal ion salvation. Both solvent separated and contact ion pairing were observed for lithium nitrate solutions, but not for the ammonium nitrate solution. In addition, two bands at 561 and 361 cm"1 are assigned to cation—solvent, Li-N, stretching vibrations. At the same time, Roberts, 33, 31, (85) studied the effect of sodium perchlorate and iodide on the hydrogen bonded structure of liquid ammonia. In addition to the expected changes IS in the N—H stretching region, a broad, weak, polarized band at m200 cm.1 was observed in the sodium perchlorate spectrum. Two bands at 325 and 450 cm.1 were observed for the sodium iodide solution. In both cases, these anion dependent bands could not be attributed to salute or solvent; hence, it was proposed that solvent separated ion pairs were responsible. Plowman and Lagowski (86) examined Raman spectra of ammonia solutions of some alkaline earth and alkali metal perchlorates and nitrates. In all cases, the perchlorate bands were unperturbed whereas the nitrate bands were split. Again, low frequency bands assigned to the symmetric stretching mode of the solvated cation 2+ 2+ were observed for Li+, Na+, Mg , Ca , Sr2+, and Ba2+ at 241, 194, 328, 266, 243 and 215 cm.1 respectively. CONCLUSIONS It thus appears that spectroscopic techniques such as nuclear magnetic resonance, infrared and Raman vibrational spectroscopy may be used to characterize ion—ion and ion-solvent interactions in solutions. The aim of such investigations would be to comprehend more fully the effect of the solvent on these interactions and then to relate these effects to physicochemical properties of the solvent. GMWERII EXPERIMENTAL PART SALTS Sodium tetraphenylborate (J. T. Baker), thiocyanate (Mallinkrodt), perchlorate (G. F. Smith), iodide and bromide (Matheson, Coleman and Bell) were of reagent grade and were used without further purification except for drying. The first two salts were dried under vacuum at 60°C for 72 hours, and the other salts were dried at 110°C for 72 hours. After drying, all salts were stored in a vacuum dessicator charged with granulated barium oxide. SOLVENTS Nitramethane (Aldrich) was fractionally distilled over granulated barium oxide followed by drying over freshly activated Linde 5A molecular sieves for 24 hours. Acetonitrile was fractionally distilled over calcium hydride after refluxing for 48 hours. Sulfolane (Shell) was purified by fractional freezing six times followed by fractional distillation over sodium hydroxide pellets. Propylene carbonate (Aldrich) was dried over Linde 4A molecular sieves followed by vacuum distillation at 40 torr. Acetone (Fisher), ethyl acetate (J. T. Baker) and tetrahydrofuran (Matheson, Coleman and Bell) were fractionally distilled over calcium sulfate (Drierite), phosphorous pentoxide and calcium hydride, respectively. Formamide (Fisher) was purified by six repeated fractional freezings. Methanol (Baker) was fractionally distilled over calcium sulfate. Dimethylformamide (Fisher) was dried over Linde 4A molecular sieves followed by vacuum distillation over phosphorous pentoxide. Tetramethylurea (Aldrich) and tetramethylguanidine (Eastman) were purified by refluxing over granulated barium oxide followed by fractional distillation under vacuum. Dimethylsulfoxide 16 17 was dried over Linde 4A molecular sieves and vacuum distilled. Ethanol (commercially available, 200 proof) was fractionally distilled over calcium hydride. Pyridine (Fisher) was refluxed over granulated barium oxide and fractionally distilled. Hexamethylphospharamide (Aldrich) was vacuum distilled over granulated barium oxide at 20 torr. Acetic acid (Baker) and formic acid (Baker) were each purified by six repeated fractional freezings. Purified solvents were stored over Linde 4A molecular sieves in brown bottles with the tops covered with cellophane wrap. Important solvent properties and solvent abbreviations to be used in this thesis are listed in Table 2. SAMPLE PREPARATION For the NMR concentration studies, stock solutions of the sodium salts, generally 0.50 M, were prepared by weighing out the desired amount of salt into a 5 ml volumetric flask and diluting to the mark with solvent. The remaining solutions were prepared by appropriate dilutions of the stock solution. The mixed solvent solutions were prepared by taring a snap cap vial, adding the desired volume of solvent A, weighing, adding the desired amount of solvent B and weighing again. Knowing these weights, the number of moles of each solvent and resultant mole fraction were calculated. Approximately 0.1171 g of sodium tetraphenylborate was weighed into a 1 ml volumetric flask and the desired solvent or solvent mixture was then added up to the mark (No.50 M_Na+). Samples for the infrared and Raman studies were prepared in 1116 Same manner. 18 «m.Hm w.mN m.wN 0.0N «m.mN an.vN o.oN H.NH o.NH H.mH w.vH H.vH N.N Hopesz Hocoo m.:cm5uso mm.e~ mam.o wo.oe omo.o oo.m~ fine 0 Ha.cm oom.o k.mm omm.o m.aoa 2mm.o mm.a aam.o No.c omm.o k.om oce.o o.mc oac.o o.me weo.o w.wm amm.o s.mm Hmm.o Beachcou OH x a- uwspoofiofla Nuflflfinflpmoomom oflouoEDHo> Hocmcum AOmZQV ovwxomHSmecuoEMQ mazev moasfixnuoemauoe flazev oeAEaenomHEESosaa Hocmguoz oowEmEpom AmIHV Caedmopwxnmhuoh oompoomaxcum o:ouoo< numv oumconumo oonAQOHm ocmfiomfism Azuav oHAeoAcoeoo< QCGSHGEOHHHZ pco>How mofluaomoom uco>aom xox .N manme 19 mm coeoaeoea. Howe: wflo< ofiswom Cflo< 0fluo0< Auzkv ocfleficeomflxguoEmoooH maom ausz msooss< m o.m a> macaem Haoaeoeu oz .m oases MN 28 00.0 00.0: 00.0: 0v.0: 00.0 00.0 00.0 0~.v 00.0 00.0 0v.5 00.0 00.0 00.0: 00.0: 00.0: 0H.H 00.0 00.0 00.0 00.0 0N.v 00.0 00.0 05.0 00.0 00.0 00.5 00.0 0H.0 00.5: 0v.0: 00.0: 00.0 00.0 0H.0 00.v 00.0 00.v 00.0 0N.0 05.0 00.0 00.0 0v.5 00.0 0N.0 00.5: 0v.0: 0H.0 00.H 00.0 00.0 00.0 00.v 0N.v 00.0 00.0 00.0 00.0 0N.0 0v.5 ofiumz 00.0 0N.0 0~.5: 0v.0: 0v.0 05.H 0N.0 0N.0 0H.0 0H.v 00.? 00.0 0v.0 05.0 00.0 0N.0 00.5 00.0 0N.0 00.5: 0v.0: 0v.0 00.H v0.0 0N.0 00.0 0H.v 0H.v 00.0 00.0 05.0 00.0 0H.0 00.5 00.0 00.0 00.5: 0v.0: 05.0 0H.N 0v.0 00.0 00.0 0N.v 00.0 00.0 05.0 00.0 00.0 0N.0 05.5 aflo< oafihom ufio< uapoo< ocfivwemawfixnuoamnpoe ovflsmuocmmozmaxguoamxo: ocwvwnxm Hocmnom ooaxomHsmHREDosAa mannaxcuoEMHuoh owMEmEHomenuoafia Hocmnuoz owMEHBHOm :MHSMOHvxnmnuoh oumuoomfixnum o:ouoo< oomconhmu oaoaxmonm ocmaomHsm mafinuflcouoo< QCNr—HOEOHHMZ Aconcaocouv m oases 30 00.0 0N.00: 00.0: 0v.5: 00.0 00.0 00.N 00.v 00.0 0N.v 00.0: 0 00.0 00.0 0v.0 00.0 00.00: 00.0: 0N.0: 00.0 00.0 00.0 00.v 00.0 0N.v 0N.0: 05.0 00.0 00.0 00.0 0v.0 0v.0 00.00: 00.0: 0N.0: 00.0: 00.0 00.0 05.0 00.0 0N.0 0v.0: 0N.v 00.0 00.0 00.0 om.w ce.e when HI 00.0: 00.0 00.0 00.0 00.0 00.v 0v.0 00.0 00.5 00.N 0N.0 00.5 00.0 sz oc.w mm.c oo.aa- oc.m- o2.a- ca.o- mo.o oe.o mm.e m.m 00.0 00.N 00.5 00.5 00.v on.w mm.c mm.ma- mc.m- om.a- mm.o- ea.o mo.o mm.e om.m ma.e 00.0 00.5 00.0 00.0 00.0 0v.0 00.00: 05.0: 00.0: 0N.0: 00.0 00.0: 0N.0 00.v 00.0 00.0 00.0 0v.0 00o< O0Euom 00o< owuoo< 0:000cm50050uoemuu05 000amnocmmocm0xapoemxoz 0:000me 00cmzum ocaxomHSoHAEHoeAa moas0xnuosmuuob o005m500005000600 00cm0uoz 0005msnom amasmonwxcmnuoh oumHoOM0xnum onouoo< opmconumu oco0xmoum o:M0ow0:0 o00Hp0couoo< ocmnposoau0z mooscfiecouv m oHoee 31 00.0: 0N.v 00.0 00.0 00.0: 00.0 0N.v 00.0 00.0: 00.N 0N.0 00.0 00.0: 00.0 00.0 0v.0 Hmmz 00.0: 0v.0 00.v 00.0 0N.0: 00.0 00.0 00.0 00.0: 00.0 00.v 00.0 ocaxomHsAerDoeAa 000EmEa00050u0600 0000:8890"— accesses fleescaocouv m oaooh 32 Figure 2 displays the data obtained in sulfolane (D = 44.0) and acetone (0 = 20.7). As in the previous case, the thiocyanate and the iodide salts exhibit marked concentration dependence attributable to contact ion pair formation whereas the perchlorate and tetraphenyl- borate salts are relatively concentration independent. Moreover, the shifts seem to be converging toward a single point at infinite dilution characteristic of free sodium ion solvated by acetone or sulfolane respectively. It is interesting to note that in formamide (D = 109.5) (Table 1) the chemical shifts of all salts used do not display concentration dependence, and the positions of the 23Na resonance for each salt are almost identical (~4.20 ppm). These data suggest that over the concentration range studied (0.01-0.50 M) contact ion pairs do not exist in these solutions. Classically, such behavior is to be expected in solvents of high dielectric constant. However, the 23Na chemical shifts in solvents of very low dielectric constant such as tetramethylguanidine (TMG) (D = 11.0), pyridine (D = 12.3) and tetrahydrofuran (THF) (D = 7.6) also show little or no concentration dependence from 0.01-0.50 M_as shown in Figures 3 and 4. It should be noted, however, that in these solvents the chemical shifts do not converge at lower concentrations as they do in solvents of medium and high dielectric constant. In solvents of low dielectric constant, the concentration of free ions is vanishingly small in the 0.01-0.50 M_range, and the predominant species must be contact ion pairs. It should be noted that in the chemical shift vs concentration plots, the lines do seem to be curving upward in the 0.01-0.l M_concentration range where one would expect 33 10 __ SULFOLANE 9 .— E} .— 7 0 6 -—J 5 _J 9 __ ACETONE E 8 _ c s v——1——-——o-——-° 0%,, G. d 7 —'i I I l l l l 0.0 0.1 0.2 0.3 0.4 0.5 CONCENTRATION (M) Figure 2. Sodium-23 Chemical Shifts of Various Sodium Salts in Sulfolane and Acetone 34 1 _ PYRIDINE CIO W e 4" 4 '1 -!:— 3“ u ”if “—U 9T* 49 BPh4 -3 '- v—r : 40 v 4 SCN -5 .0. .37 .— 2;- e a. -9 - ° I 0. <1 -6 __ TETRAMETHYLGUANIDINE K n o C —-d CIO4 -{3 .0 oo\tt\<> . 0 a 9 e SCN '10 -‘\ ‘U‘ 0 J3 BPh4 -1;Z .— -14 -°\°\°\ ° . . e I I I l I I T 0.0 0.1 0.2 0.3 0.4 0.5 CONCENTRATION (M) Figure 3. Sodium-23 Chemical Shifts of Various Sodium Salts in Pyridine and 1,1,3,3-Tetramethy1guanidine 35 10 '— 8 _ 9-9 Jr 0 9 00;) 00,, W 9* 17—4 BPh4 6 .. 4 _ 2;. _ 0—-0 SCN (1. 2 0. fl 0 __, -2 .1 -4 .— -6 —\9\0 I ‘3 I I I T l 0.0 0.1 0.2 0.3 0.4 0. 5 CONCENTRATION (M) Figure 4. Sodium-23 Chemical Shifts of Various Sodium Salts in Tetrahydrofuran 36 ion pair dissociation to become observable. The present detection limit of 0.01 M solutions prevents investigations at lower concentrations. The conductance of Nal solutions in TMG and pyridine was determined 5-5 X 10‘1 M concentration range. The data are shown in the l x 10‘ in Figures 5 and 6, respectively. It is seen that ion pair dissociation is negligible in solutions with concentrations above 10.2 M, The conductance data were analyzed by the Fuoss-Shedlovsky technique (92). For TMG, A0 = 51.6 and the ion pair dissociation constant, Kd = 6.2 x 10‘5. In pyridine, A0 = 73.4 and Kd = 3.91 x 10‘4. The latter ; I value agrees with Kd = 3.7 X 10-4 as determined by Burgess and Kraus (93). No literature data seem to be available for the conductance of sodium iodide in TMG. A literature search for conductance studies of the other salts in these three solvents revealed only the following data for sodium tetraphenylborate in THF; A0 = 88.5, Kd= 8.52 X 10"5 (94,95). Hence, the conductance data indicate that for the salts studied in these low dielectric solvents, ion pairing phenomena prevail over the concentration range examined in the NMR study. Results obtained in hydrogen bonding solvents such as methanol (0 = 32.7) and ethanol (D = 24.5) show one significant difference as compared to the other solvents studied thus far. As shown in Figure 7, the 23Na resonance observed for NaClO4 in both Solvents and NaBPh4 in ethanol shifts upfield with increasing concentration, while in the case of NaI and NaSCN the cOrrespoNding shifts are dawnfield. Upfield shifts have previously been reported for aqueous alkali metal nitrate solutions (16). With increasing concentration, the alkali metal resonances of the nitrate solutions shifted upfield while those 37 50'- 40—? 30" -91r-c-t-—$ Atom/Is"CMZEOUIM") 20'— 10— ' 3’ 313 953 ::: e 3 3 =$-o a a a 0‘ l I l l I l I 00 01 a2 03 a4 05 06 D7 {CU2 ~ Figure 5. Conductance Curve for Sodium Iodide in l,l,3,3-Tetramethyl- guanidine 38 70-1 60- 0" O l w—cfl“ A (OH MS'ICMzEQUIV.'1) 00 «b o O l : 200 10"“ T l l l l I I l 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 CI/ 2 Figure 6. Conductance Curve for Sodium Iodide in Pyridine 39 METHANOL 004 ‘4 .7:y——“————‘> A o-——o 09:14 L) Q U G U W SCN . Br,I ETHANOL 004 BPh4 '0 SCN '1 I l l l i 0.0 0.1 0.2 0.3 0.4 0.5 CONCENTRATION (M) Sodium—23 Chemical Shifts Of Various Sodium Salts in Figure 7. Methanol and Ethanol. 40 in alkali halide solutions shifted dawnfield. Similar behavior has recently been observed by Van Geet, again, in aqueous solution (35). While both upfield and dawnfield shifts with increasing salt concentration are due to increasing cation-anion interaction, it has been postulated that upfield shifts occur when a water molecule in the cation salvation shell is replaced by an anion, which results in a decreased electron density around the alkali cation. Although it is possible that the same explanation applies to the upfield shifts observed in methanol and ethanol solutions, a detailed NMR and vibrational study of systems exhibiting upfield shifts was pursued. This study is presented in Chapter V of this thesis. 'To this point, we have been examining contact ion pair formation in several nonaqueous solvents and rationalizing the data with respect to the dielectric constant of the solvent. Consider the results obtained in propylene carbonate (PC) (D = 65.0) and dimethylsulfoxide (DMSO) (0 = 46.7), as shown on Figure 8. In PC, contact ion pairing occurs in iodide and thiocyanate solutions, whereas perchlorate and tetraphenylborate solutions are composed of free solvated iohs and/or solvent separated ion pairs. In DMSO, a solvent of’lggg£_dielectric constant, not only is there no contact ion pairing in the perchlorate and the tetraphenylborate solutions, but there is no contact ion pairing in the iodide or the thiocyanate solutions as well. It appears that, to a large degree, ion pair formation is closely related to the bulk solvent dielectric constant; however, these studies show that other factor(s) may be important. After defining a quantity which describes solvent donicity on a molecular basis (as opposed to 41 10 __1 PROPYLENE CARBONATE G O'——") C|O4 _ 0 o C} -—— BPh4 9 -—I 8 .— '7 .— 5? I g] 6 _4 ° 0 0a SCN ‘I — DIMETHYLSULFOXIDE BPh4 O _ .1--. ‘— I"! ‘)‘——() I SCN -1 __ Br -;Z._a I I l l I N 0.0 0.1 0.2 0.3 0.4 0.5 CONCENTRATION (M) Figure 8. Sodium-23 Chemical Shifts of Various Sodium Salts in Propylene Carbonate and Dimethylsulfoxide 42 bulk preperties) we will investigate the interplay of solvent dielectric constant and donicity on contact ion pairing phenomena. The frequency of the 23Na resonance of sodium ion in solution is affected primarily by the nearest neighbors of the ion, hence if the effects of cation—anion interactions were removed from the system under investigation, the observed resonance would indicate solely the interaction of the ion with the solvent. Hence, the variation of the 23Na chemical shift with solvent may be explained by the ability of the solvent to alter the electronic environment about the sodium-23 nucleus, which should be related to the donor ability of the solvent. An empirical approach to this donation or camplexing ability of the solvent is given by Gutmann's donor numbers (96). These are simply the enthalpies of complex formation between the given solvent and antimony pentachloride in 1,2-dichloroethane solution. As shown by Gutmann, the donor numbers may be quite useful in predicting the behavior of nonaqueous systems. It is reasonable to expect, therefore, that there might be a relationship between the donor numbers and the relative chemical shifts for the solvated sodium ion provided that the chemical shift is unperturbed by cation—anion interactions. A plot of the infinite dilution 23Na chemical shift in a given solvent versus the donor number of that solvent, shown in Figure 9, reveals that this is indeed the case. The linearity of the plot shows that the relative 23Na chemical shifts yield useful information on the solvating abilities of the solvents at least vis a vis the sodium ion. 43 A 53 I K APPM '4 I | l l 0 10 20 30 40 DONOR NUMBER Figure 9. Plot of Infinite Dilution Sodium-23 Chemical Shifts versus the Donor Number of the Solvent: (1) Nitramethane, (2) Acetonitrile, (3) Sulfolane, (4) Propylene Carbonate, (S) Acetone, (6) Bthylacetate, (7) Tetrahydrofuran, (8) Dimethylformamide, (9) Tetramethylurea, (10) Dimethyl- sulfoxide, (11) Water, (12) Pyridine, (l3) Hexamethyl- phospharamide, (14) Formamide, (15) Methanol and (16) Ethanol 44 The plot of infinite dilution chemical shifts as a function of donor numbers was fitted by linear least squares analysis to yield the following equation: chemical shift = (-0.5074)(DN) + 16.6105 (3) The donor numbers of methanol, ethanol and formamide have not been determined experimentally but from the above equation they can be predicted to be 25, 32 and 24 respectively, again provided that the chemical shifts are unperturbed by ion-ion interactions. It should be noted that, while the salvation process involves electrostatic ion-dipole interactions, donor numbers represent the enthalpy of formation of a covalent bond between a given solvent and antimony pentachloride. A priori it seems, therefore, that a parallelism between the solvent's ability to solvate alkali metal cations and its tendency to form covalently bonded complexes with SbCl5 should not be expected. As Gutmann points out, however, the fact that such parallelism exists suggests that some covalent interaction is involved in the salvation of sodium ion by donor solvents and that the extent of covalency in the salvation bonds increases with the solvent donicity (97). On the other hand, this parallelism may suggest that some electrostatic interaction is involved in the solvent-antimony pentachloride bonding. The fact that donor properties of solvents can strongly influence ion pair dissociation has been shown by Gutmann (97). The data reported in this thesis suggest that both solvent dielectric constant and donicity influence ion pair dissociation, as summarized in Figure 10. 45 10's 9 .. 3 "'I SULFOLANE 7 " o 6 .. E 5 ._ 0.. a. FORMAMIDE <1 03—00 0 (I C O I I I I I 0.0 0.1 0.2 0.3 0.4 0.5 CONCENTRATION (M) Figure 10. Sodium-23 Chemical Shifts of Sodium Iodide in Sulfolane, Formamide, Acetone and Dimethylsulfoxide 46 For example, in formamide solutions (D = 109.5, DN = 24) the formation of contact ion pairs in the 0.01-0.50 M_concentration range would not be expected due to the high dielectric constant and medium donor number of the solvent. The data seem to confirm this conclusion. The comparison of dimethylsulfbxide (D = 46.7, DN = 29.8), sulfolane (D = 44.0, DN = 14.8) and acetone (D = 20.7, DN = 17) shows the importance of the donicity of the solvent in contact ion pair formation. While dimethylsulfoxide and sulfolane have nearly identical dielectric constants, the former is a much better solvating agent, hence contact ion pair fOrmation is expected to be more probable in the latter solvent. Indeed, the 23Na chemical shift fbr NaI shows no concentration dependence in dimethylsulfoxide but a marked concentration dependence in sulfolane. 0n the other hand, acetone and sulfolane have nearly the same donicity, but the latter has a much higher dielectric constant. Qualitative indications from the data reveal the tendency toward ion pair formation in acetone is much higher than in sulfolane. Additional evidence has been obtained in propylene carbonate solutions (D = 65.0, DN = 15.1). As shown in Figure 8, the 23Na chemical shifts for sodium iodide solutions are stronger concentration dependent, which indicates that despite the high dielectric constant, the formation of contact ion pairs occurs to a considerable extent. While to this point only qualitative indications of ion pair formation have been obtained from 23Na chemical shifts, it seems reasonable to assume that these data can be used for the determination of ion pair formation constants. Since only one signal is observed 47 for the 23Na resonance in solution, it may be assumed that exchange is rapid compared to the NMR time scale; hence, a population weighted average shift is observed and is described by equation (4): sobs = XF6F + XCGC (4) bs is the observed chemical shift in ppm, 6F is the chemical shift characteristic of free solvated sodium ion in a given solvent where 5 0 in ppm, 6C is the chemical shift characteristic of the contact ion pair in that solvent in ppm, and X and XC are the fractions of free F and complexed sodium ion respectively. Now, sobs = XFGF + (1 - XF)6c (5) Rearranging, Gobs = XF(aom xn madam Esflpom msoflum> How mucmumcou :ofiumahom yawn :oH .v manna CHAPTER IV STUDIES OF PREFERENTIAL SOLVATION OF THE SODIUM ION IN MIXED SOLVENTS BY SODIUM-23 NMR INTRODUCTION In the preceding chapter, we saw that NMR spectroscopy may serve as a valuable tool for the elucidation of the structure of electrolyte solutions. The variation of the chemical shift as a function of solvent, anion of the sodium salt and concentration allows one to differentiate contact ion pairs versus free solvated ions or solvent separated ion pairs. More importantly, the magnitude and direction of the 23Na chemical shift varies linearly with the donicity of the solvent while the resonance shifts further downfield as solvent donicity increases. These studies were extended to binary solvent mixtures where the variation of the 23Na resonance as a function of solvent composition was monitored and interpreted in terms of preferential solvation. It is further hoped to rationalize preferential solvation as a manifestation of solvent donicity. When an electrolyte is dissolved in a binary solvent mixture, the first solvation shell of the cation need not have the same composition as the bulk solvent composition; it may preferentially contain one solvent over the other. If the variation of the cation resonance is monitored as a function of mole fraction of one solvent in a binary mixture, generally, a smooth transition of the resonance from its value in one pure solvent to the other is observed where the rate of transition is dependent on the relative solvating abilities of the two solvents in the given mixture. At some point in the study, the chemical shift has progressed one half the way from one limiting value to the other. This point has been defined as the equisolvation 51 52 or isosolvation point, where there is equal competition of each of the solvent components for the cation. In this chapter, the isosolvation point will be employed as a measure of preferential solvation. Frankel, etc 31: (99) examined preferential solvation of Co3+ ion (Co(acac)3) in chloroform-carbon tetrachloride mixtures by 59Co NMR. Likewise, preferential solvation of Cr3+ ion (Cr(acac)3) in the same solvent system was examined by observing the effect of the paramagnetic Cr3+ ion on the transverse relaxation time (T2) of the solvent nuclei (100). These same workers also studied 35Cl chemical shifts in dimethylsulfoxide-water and acetonitrile-water mixtures, but were not able to draw any conclusions concerning preferential solvation (19). Bloor and Kidd (33) examined 23Na resonance of sodium iodide solutions in acetonitrile-water mixtures and concluded that Na+ ion was preferentially solvated by water. RESULTS AND DISCUSSION The chemical shift of the sodium-23 nucleus as a function of solvent composition in binary solvent mixtures of nitromethane with dimethylsulfoxide, pyridine, tetramethylurea, hexamethylphosphoramide and acetonitrile; acetonitrile with pyridine, dimethylsulfoxide, tetramethylurea and hexamethylphosphoramide; tetramethylurea with pyridine and dimethylsulfoxide; hexamethylphosphoramide with tetramethylurea, dimethylsulfoxide and pyridine, and dimethylsulfoxide with pyridine using sodium tetraphenylborate as the solute are presented in Table 5 and illustrated in Figures 11-14. 53 mN.m I 000.H m~.m I 005.0 00.H I 000.H mN.n 000.H 00.m I w0m.0 0N.H I 0mw.0 0v.h N00.0 0m.m I m0¢.0 0m.0 I 000.0 mw.n H00.0 mm.H I m0m.0 00.0 Nov.0 00.0 H0m.0 m©.0 I 0m~.0 mw.H 00m.0 0N.0 0Hm.0 mm.~ m0~.0 0N.v 00H.0 m.0H v0N.0 0N.m mmo.0 mm.m HmH.0 0.HH ovH.0 mn.m 000.0 0m.n 000.0 N.NH 000.0 v.0H HN0.0 0.0a Nm0.0 o.mH wv0.0 m.mH 000.0 n.va 000.0 N.mH 000.0 200< <02: 02 Emm< mzHQHm>m m: Emmq qumHHZOHmu< 02 mz0 mzHom thcfim How cofiuwmomsou uco>Hom mo coHuUCSm < mm mocmcomom MNIESflvom map mo :ofiumfihm> .m oHan 54 mo.HI m5.NI 00.NI 0N.0I 00.0I mm.mI 0m.mI 00.0I 00.0I 00.0I mm.mI 200d 000.H 500.0 000.0 H05.0 «00.0 H00.0 00v.0 00N.0 mom.0 00H.0 000.0 mzHQHm>0 02 <0EIHMZHQH0>Q 00.N 0N.0 00.0I mm.HI 05.HI mv.NI 05.NI 0H.mI 0v.mI mv.mu mm.mI 200< 000.H v00.0 505.0 005.0 m0m.0 00v.0 00v.0 v0m.0 v0H.0 500.0 000.0 02% m: 0N.0 mm.HI 00.NI 0v.mI 00.NI 00.mI mH.mI 0~.mI mm.mI m©.mI mm.mI 200< 0 000.~ N00.0 N55.0 N05.0 N00.0 000.0 v0v.0 50N.0 00N.0 mHH.0 000.0 mZHQHm>0 02 0N.N mN.N 00.N 00.N 00.0 00.0 00.0 mo.v 00.0 05.0 Emm< 000.H «00.0 m0m.0 000.0 H00.0 00H.0 0vH.0 000.0 0v0.0 000.0 32% 02 qumHHZOHmum mAHmHHZOHmum 02 0N.0 00.0 05.0 00.0 00.H 00.H 00.H 00.N 0N.0 05.0 200< 000.~ 0H0.0 000.0 000.0 00N.0 va.0 m0~.0 00H.0 vHH.0 000.0 0020 m: maHmBHZOHm00uomEQ 0N.0 00.H 05.0 0H.H 00.H 00.? 00.0 00.5 v.HH 0.0M 2004 000.H 005.0 000.0 N0¢.0 00N.0 NOH.0 V50.0 mmo.0 mm0.0 000.0 0020 02 mz Py > ACN >>> CHSNO2 which is to be expected as shown by the studies of donicity in Chapter III. The only exception seems to be pyridine. With a donor number of 33.0, pyridine should be a stronger donor than DMSO (DN = 29.8) and TMU (DN = 28.9) but weaker than HMPA (DN = 38.0). Finally, it is interesting to note that HMPA, DMSO and TMU seem to exhibit the same relative donor abilities in nitromethane although their respective donor numbers are widely differing. Hence the donicity of these solvents in nitromethane is "leveled". 63 Table 6. Isosolvation Points for Solvation of the Sodium Ion in Binary Solvent Mixtures Binary Solvent System DMSO:Nitromethane 0.05 PyridinezNitromethane 0.12 TMUzNitromethane 0.06 HMPAzNitromethane 0.05 Acetonitrile:Nitromethane 0.15 TMU:HMPA 0.23 DMSO:HMPA 0.15 PyridinezHMPA 0.10 Pyridine:Acetonitrile 0.29 DMSO:Acetonitrile 0.10 TMU:Acetonitrile 0.11 HMPAzAcetonitrile 0.06 Pyridine:TMU 0.16 DMSO:TMU 0.39 DMSO:Pyridine 0.10 MP MF MP MP MP MF MP MP ME ME Isosolvation Point DMSO Pyridine TMU HMPA Acetonitrile HMPA HMPA HMPA Pyridine DMSO TMU HMPA MP MP MF TMU DMSO DMSO < h'mV-V‘ '1-_:w I._.- 64 Acetonitrile is a solvent of medium donicity (DN = 15.0) and as such should be more competitive with the preceeding solvent series for Na+ ion than was nitromethane. Binary solvent mixtures of acetonitrile with pyridine, dimethylsulfoxide, tetramethylurea, hexamethylphosphoramide and nitromethane exhibit isosolvation points of 0.29, 0.10, 0.11, 0.06 and 0.85 mole fraction of the latter respectively, as shown in Figure 12. Again, the strong donor solvents HMPA, DMSO and TMU are "leveled” but we now seen an even greater differentiation of pyridine from these solvents. In acetonitrile, the relative order of donor ability is HMPA 3 DMSO 3 TMU > Py > ACN >> CH3N02. The "repression” of the donicity of pyridine in binary mixed solvents was surprising. The data seem to indicate that DMSO and TMU are better solvating agents for the sodium ion than pyridine, which is contrary to the 23Na NMR chemical shift data and to the solvents' donor numbers. To investigate further this unexpected donor repression of pyridine relative to other high donor solvents, solvent mixtures of pyridine with hexamethylphosphoramide, tetramethylurea and dimethylsulfoxide were examined. The 23Na nmr data revealed isosolvation points of 0.10, 0.16 and 0.10 mole fraction of the latter indicative of strong preferential solvation by that solvent. Recalling the donor numbers of 38.0, 29.8, 28.9, and 33.0 for HMPA, DMSO, TMU, and pyridine respectively, we would expect small preferential solvation of the Na+ ion by HMPA with respect to pyridine, but very little or no preferential solvation of the Na+ ion by DMSO L 65 or TMU relative to pyridine since these three solvents have similar donor numbers. Instead, strong preferential solvation of the Na+ ion is noted for DMSO and TMU versus pyridine. It was decided to examine the mixed solvent system DMSO-pyridine in greater detail using vibrational spectroscopy to gain further insight into either the enhanced donicity of DMSO or the repressed donicity of pyridine in their mixtures. It has been shown in previous publications that the vibration of the sodium ion in the DMSO solvation shell is observed at 200 cm-1 (53), while in the pyridine solvation shell the corresponding frequency is 180 cm-1 (58). The position of the solvation band in DMSO-pyridine mixtures was determined and the results are summarized in Table 7 and are shown in Figure 15. It is seen that even the addition of small amounts of DMSO to pyridine results in a nearly complete shift of the solvation band frequency to that corresponding to pure DMSO. In fact, the rate of change of the band frequency with composition strongly parallels the corresponding change in the 23Na chemical shift. These results seem to confirm strong preferential solvation by DMSO contrary to "donicity" values. A possible explanation of this apparent anomaly may be in the different structures of the two solvents. As evidenced by its Trouton constant of 29.6 cal deg-1 mole-l, DMSO is a highly associated liquid with a chainlike structure consisting of aligned S-O dipoles (102). Even in dilute benzene (102) and carbon tetrachloride solutions (103) the molecules are said to be associated into a cyclic dimer. It seems reasonable, therefore, to assume that when a sodium salt is introduced into neat DMSO, a considerable amount of energy Table 7. 66 Variation in the Frequency of the Sodium Ion Solvation Band for 0.50 M_NaBPh Mixtur MOLE FRACTION O 0 0. 85 DMSO .000 .050 099 .151 .198 .300 .400 .501 .598 .703 .800 .946 .000 4 Solutions in DMSO-Pyridine Solvent FREQUENCY (cm-1) 178 :.2 184 188 191 195 198 200 201 199 199 201 200 201 I'm x—an- . II c..- 67 I. monsuxfiz ocflvwnx0IonfixomHamfixnuosfio a“ e200mz : 00.0 mo teem nowum>aom gem samvom one mo 50co30090 059 we cowumwnm> .mH onswfim OmED 20.....04m—n. MACS— O.— 0.0 0.0 5.0 ed 0.0 .vd md Nd .5 Cd _ _ . PI PI — _ — — — _ — myFP meowusaow I 00— I In 0 ._ r O 92 (kwm ACNE-10038:! I we * v I com I mow “— —-°_ 68 must be expended to break up the structure of the solvent before solvation of the cation can occur. 0n the other hand, pyridine is a relatively unstructured polar liquid with a dipole moment of 2.2 D. The introduction of even small amounts of pyridine into neat DMSO may result in the break up of the polymeric structure of the latter solvent via a dipole interaction. To investigate this possibility, the S-O stretching frequency of DMSO was monitored in neat DMSO and in DMSO-pyridine mixtures. The variation of the S-O stretching frequency with solvent composition would be indicative of DMSO-pyridine interactions causing structural changes in the medium. The data are presented in Table 8 and Figure 16. It is seen that the addition of small amounts of pyridine to DMSO results in a sharp increase in the vS-O frequency to N 0.05 mole fraction pyridine (0.95 mole fraction DMSO). In the region 0.10-0.80 mole fraction pyridine, the v frequency gradually S-O increases. Above 0.80 mole fraction pyridine, the increase in the VS-O frequency becomes steep again. These data then do reveal a strong solvent-solvent interaction causing extensive disruption of DMSO structure by pyridine. A recent study of the Brillouin scattering by pyridine-DMSO mixtures confirms the conclusions reached by examining nmr and vibrational data (112). Although no specific vibrational study of the TMU-pyridine system was undertaken here, it should be noted that the enhanced donicity of TMU relative to pyridine may also be rationalized in terms of structure breaking by pyridine. As Luttringhaus (104) points out, TMU is structured by strong dipole association as is 69 Table 8. Variation of the S-O Stretching Frequency of DMSO for DMSO-Pyridine Solvent Mixtures MOLE FRACTION FREQUENCY DMSO ' (cm-1) 0.050 1068 1 1 0.100 1064 0.150 1063 0.198 1063 0.300 1062 0.400 1061 0.500 1061 0.598 1059 0.700 1057 0.800 1057 0.946 . 1056 1.000 1049 70 moHSHxfiz onwvflax0Iowfixowfismchuosmo new xocosvoum mcwgouonum 0-0 0:9 mo :ofiumfiua> .oH 093000 Oman—0 20:049....— MAO—2 0.5 0.0 0.0 5pc 9.0 0.0 Jo 0.0 N._0 5 0.0 _ _. _ _ _ 32 I So. 14‘ H 1.. I 89 m a N 3 I 82 M 3 w. I mac. 1 r 22 71 DMSO. Beguin and Gunthard (105) showed the C=0 stretching frequency to be strongly solvent dependent as is the S-O stretching frequency in DMSO. Hence, a strong TMU-pyridine dipole-dipole interaction may account for the enhanced donicity of TMU by causing extensive disruption of TMU structure by pyridine. Finally, we wished to differentiate the three solvents that were effectively leveled in nitromethane and acetonitrile.« For binary . solvent mixtures of hexamethylphosphoramide with tetramethylurea and dimethylsulfoxide, the isosolvation points were 0.23 and 0.15 mole fraction HMPA as shown in Figure 13. An isosolvation point of 0.39 mole fraction DMSO was noted for the dimethylsulfoxide-tetramethylurea system shown in Figure 14. These data imply that the relative order j of solvating ability is HMPA > DMSO 3 TMU where DMSO is just slightly stronger than TMU. To this point preferential solvation has been discussed in a qualitative sense as reflected by the isosolvation point. Covington, 33, El: (106,107), points out that if we make a few reasonable physical assumptions, we may develop a quantitative model for competitive solvation. In the above papers, he presents a thermodynamic derivation of preferential solvation that allows the calculation of equilibrium constants and the changes in free energy as the solvation shell of an ion X is progressively changed from n molecules of solvent W to n molecules of solvent P. The remainder of this chapter will be devoted to a discussion of the physical assumptions underlying Covington's treatment followed by an application to the data presented in this chapter. 72 Let us assume that initially ion X is solvated by four molecules of the solvent W. As the second solvent P is introduced into the system, there is a step—wise replacement of W by P in the inner solvation shell. The series of equations can be expressed as, K K 1 2 XW4 + pP + (w-4)W-——-—+ XW3P + (p-1)P + (w-3)W'-——+ K3 K4 -——+ ———+ XWZP2 + (p-2)P + (w-2)W'+-- XWP3 + (p-3)P + (w-1)W'+-———— XP4 + (p-4)P + wW , (17) Considering the first step as an example, we can write o__ =0 _o AG1 — RTJLnK1 “stp “XW4 + u; - u; (18) At this point, we assume that the chemical potential, u°, is made up of a number of contributions, for example 0 0 int 0 hyd 0 int 0 elec ochem p = u + u = u + u + u p (19) XP4 X XP4 X XP4 XP4 . Oint o o I o o hyd o where “X is the contribution from the bare ion, ”XP is the 4 contribution from the solvation shell and bulk solvent, which is then °elec XP 4 . . ochem . . . contribution,uXP , dependent only on the compOSition of the solvation 4 split up into an electrostatic contribution, u and a chemical shell. If we insert equation (19) into equation (18), we obtain 0 _ 0 Chem 0 Chem 0 o AG1 ‘ "stp ' uxw4 + “w + “P (20) 73 o elec where the 0 terms cancel if we assume that the radii of the two solvated ions, XW and XW P are equal in the Born sense. This 4 3 assumption requires that the solvent system be isodielectric. However, the treatment may be extended to non-isodielectric systems if the electronic term can be calculated from the Born treatment. Another assumption concerns the nature of the intrinsic chemical shift of each species. If 6? is the shift in the resonance of X from pure W to pure P, then Covington assumes that wa = 0; 4 XW3P = 1/4 6P; 6XW2P2 = 1/2 6P; 6 = 3/4 6?; oxp4 = 0?. the intrinsic shifts of the various solvated species are assumed to 6 Hence, XWP3 be proportional to the amount of P which they contain. Now, there seems to be no physical evidence for this assumption. Recent 13C NMR studies for a series of halomethanes (CH4_nXn where X = Cl, Br, I) reveal that as the protons are replaced by chlorine, the chemical shift varies proportionally, however, this is not the case when protons are replaced by bromine or iodine (110). Lauterbur (109) observed that in some cases additive shift changes are produced by interchanging ha10gens in these polyhalomethanes. For example, successive replacement of chlorine atoms by bromine atoms in carbon tetrachloride causes linear variations in the 13C chemical shifts. The intrinsic shift assumption implies that each P molecule will interact equally and independently with X. That is, once a P molecule replaces a W, its interaction with X will be unchanged as the next P molecule solvates X, etc. It should be pointed out that the intrinsic shift assumption must be made so that the number of unknowns in the final equation is reduced to one. Finally, it is 74 interesting to note that it is this assumption that underlies the isosolvation point, hence, the discussion contained herein also applies to this concept. The next assumption requires that the individual equilibrium constants, K1, are related solely by statistical requirements, that is, 1/4 /4. K' = K (21) 1 ‘ (K1K2K3K4) K = 4K', K2 = 3/2 K', K 1 = 2/3 K', K 3 4 = 1/4 K' (22) This implies that the equilibrium constants are solvent independent. If there indeed is some degree of preferential solvation, then these equilibrium constants should be solvent dependent. The final equation in this treatment allows calculation of Klln as fellows, %=%“(1*—‘1'T P Kl/n.(_P_ aw) ) (23) By plotting 1/6 vs Xw/XP, we obtain 1/6P from the intercept and Kl/n from the slope. Hence, we must assume that solutions are ideal. Then, we may accordingly substitute mole fractions for activities to facilitate the calculations. However, our solutions are not ideal at all. The solvation of the sodium ion, which is essentially an ion-dipole interaction, contributes to non-ideality. To expect no solvent-solvent interaction is unrealistic, as we have shown marked solvent interaction in dimethylsulfoxide-pyridine mixtures earlier in this chapter. 75 Finally, Covington concludes that since the plot of equation (23) o elec is linear as predicted, the assumptions that u terms cancel in equation (20) and that assuming the individual equilibrium conStants are solvent independent are acceptable. We treated our data according to equation (23) and calculated the l/n geometric equilibrium constant, K , and the corresponding free 0 P.S. Table 9. It is interesting to note that a plot of 1/6 vs XB/XA did energy of preferential solvation, AG /n, which are shown in in fact yield straight lines in 2220,9353» except for the solvent system acetonitrile-pyridine, which displayed a small degree of curvature. The lines were treated by a linear least squares procedure from which we obtained the values of Kl/n. It is particularly interesting to note that such a plot for the data in DMSO-pyridine mixtures, where there exists marked solvent-solvent interactions, yields a straight line as shown in Figure 17. It must be recalled that this treatment assumes that the solvation number, n, is the same in both solvents in the given binary mixture. We noted a small amount of curvature in the plot for acetonitrile- pyridine, which may be a result of n not being the same on this case. Langford (19,108) noted curvature in similar plots, which he ascribed to non-equal solvation numbers. If the values for Kl/n are meaningful (they are certainly as valid as the isosolvation point model in that both embody the same assumptions), we may again comment on relative donor strengths. We may note the same trends in nitromethane and acetonitrile as earlier mentioned, except that in the former, TMU appears to be much more competitive than either DMSO or HMPA. Another interesting observation 76 1.0 ‘— 0.9 - 0.8 - \ 0.7-1 0.6 - 0.5 — 0-4 I I I I I 0 .- 2 4 6 8 10 XPYRIDINE/XDMSO Figure 17. Covington Plot for the Binary Solvent System Dimethylsulfoxide-Pyridine 77 Table 9. Covington Treatment of Preferential Solvation Binary Solvent System Kl/n DMSO:Nitromethane 14.4 PyridinezNitromethane 7.44 TMU:Nitromethane 33.8 HMPAzNitromethane 13.2 Acetonitrile:Nitromethane 4.48 TMU:HMPA 0.0669 DMSO:HMPA 0.166 PyridinezHMPA 0.1245 PyridinezAcetonitrile 2.93 DMSO:Acetonitrile 12.7 TMU:Acetonitrile 7.01 HMPAzAcetonitrile 62.9 PyridinezTMU 0.297 DMSO:TMU 0.92 DMSO:Pyridine 17.2 o "AGP.s./" (kJ mole-1) 6.61 4.97 8.72 6.39 3.72 -6.70 -4.45 -5.16 2.66 6.30 4.82 10.3 -3.01 -0.206 7.04 78 is that Kl/ n for TMU-DMSO mixtures seems to favor TMU as the stronger donor whereas the location of the isosolvation point seems to favor DMSO. Finally, it is interesting to note that the chemical shift vs mole fraction plot for this system is almost linear, which indicates little preferential solvation, and that Covington's 065 S /n is close to zero, indicative of no preferential solvation. CONCLUSIONS The location of the isosolvation point for a given binary solvent system gives a qualitative measure of the solvation abilities of the solvents in the given mixture. In order to extend the comparison to variations in isosolvation values to a series of binary mixtures, more knowledge is needed on the structures of the individual solvents and the changes in this structure on formation of liquid mixtures. These results emphasize the danger of extrapolating the behavior and properties of pure solvents to their properties in solvent mixtures. Covington's quantitative approach appears successful despite the number of physical assumptions, some of which are not reasonable (intrinsic shifts, solution ideality, etc.). This observation has two implications. First, as Covington states, "The treatment . . . although not completely rigorous, assists, we believe in understanding the present approach to preferential solvation . . .". 0r second, this treatment is not sensitive enough; that is, the physical assumptions are not as crucial as they appear. Since the NMR technique reflects the influence of the first solvation sphere on the ion of interest, perhaps the non-ideality of the bulk solvent mixture outside this sphere has small influence on the ion. Moreover, perhaps 79 the ion-dipole solvation phenomenon does not cause significant deviations from ideal behavior. CHAPTER V A SPECTROSCOPIC STUDY OF CONCENTRATED SOLUTIONS OF SODIUM PERCHLORATE - THE NATURE OF THE UPFIELD CHEMICAL SHIFT INTRODUCTION Generally we noted that the 23Na resonance in solutions of sodium perchlorate and tetraphenylborate exhibited no concentration dependence over the concentration range 0.01-0.50 M5 whereas, the corresponding ' shifts for the thiocyanate, bromide and iodide salts exhibited marked concentration dependence, shifting downfield with increasing concentration. In the latter case, these downfield shifts were attributed to contact ion pairing, while in the former case, the static shift was proposed to be indicative of free ions or solvent separated ion pairs. These assignments seem reasonable as it would be easier for the smaller bromide, iodide and thiocyanate anions to be part of a solvation sphere as opposed to the larger, bulky perchlorate and tetraphenylborate anions. It was pointed out before that small upfield chemical shifts of the 23Na resonance were observed for sodium tetraphenylborate and perchlorate solutions in methanol and ethanol from 0.01-0.50 M, Reexamining the data and figures in Chapter III, subtle upfield shifts of the 23Na resonance are noted for sodium perchlorate solutions in tetramethylurea, acetone, tetrahydrofuran and pyridine to name a few. The problems that confront us are: (1) What type of species is responsible for the upfield shift? (2) Are the occurence and magnitude of the upfield shift related to some solvent property or is it characteristic of the electrolyte? (3) How do these upfield shifts vary with concentration over a wider range (0.01 Mfsaturation)? To answer these questions, detailed 23Na, 1H and 3501 NMR, Raman and infrared measurements were made of sodium perchlorate solutions in several nonaqueous media and are the subject of this chapter. 80 81 RESULTS AND DISCUSSION The chemical shifts of the 23Na resonance for sodium perchlorate solutions in several solvents are presented in Table 10 and diSplayed in Figure 18. With two exceptions, the range of the upfield chemical shifts in a given solvent is large (:_1 ppm), indicative of significant changes in the environment of the sodium ion. In formic acid, water, methanol and ethanol the resonance shifts upfield linearly with concentration, whereas, non-linear shifts are noted in the other solvents. Upfield shifts of alkali metal ion resonance with increasing concentration have been noted for aqueous alkali metal nitrate solutions by Deverell and Richards (16), for aqueous solutions of sodium perchlorate by Van Geet (35), and for aqueous solutions of potassium dichromate, nitrate, chromate and sulfate by Bloor and Kidd (27). Since the chemical shift results from overlap of the outer electron shells of the ions, these authors conclude that when water in the first solvation sphere is replaced by some oxyanion, the oxyanion overlaps less effectively with the cation orbitals than water, resulting in decreased electron density about the cation nucleus and hence, an upfield shift in the cation resonance. An alternative explanation is that collisional ion pairs, if formed, remain solvent separated, and the alkali metal ion is directly surrounded only by water molecules. The shift then reflects a change in the "bond" between the cation and water as the salt concentration increases. This long range shift would be analogous to proton chemical shifts in organic compounds. Since 82 00.0 0v.0 00.0 0H.0 0H.0 00.0 00.0 05.0 05.0 00.0 0v.0 00.0 flamov owfirm 00.N 00.H 00.H 0v.~ 0N.~ 00.H 00.0 00.0 0v.0 0N.0 0H.0 00.0 A20 0:00 mZOHmu< wave: msoosvmcoz :« meowusfiom oumHoHcoho0 550000 How mumflsm Hmofiaecu «2 00.0 0v.0 0N.0 00.0 00.5 00.5 0v.5 0N.5 00.0 fleoov umfigm mqmthZOHmu< 00.N m5.H 00.H 0N.H 00.H 05.0 00.0 0N.0 0H.0 0:0 ocou mN.m o~.m o~.m oo.m OH.m oo.m mw.w ma.w oo.m mm.w 05000 om firm 00.” 00.0 00.0 05.0 00.0 00.0 00.0 00.0 0N.0 0H.0 fizv ocou z<030000>m .mH ohsmfim 0000 0000 0000 000 000 >00 econ same 000 350a 000 5x0 :000000 0:00 500 up» 500 . \\\p//r 00.00000 00.00000 00.0000< 000~< 00.0000< 00.0000< -0000111 \\\2 w> o: 0: n: v: N: o 0000 000 000 0 000 3000 econ 350 ppm 500 b 00.00000 0000 0000 -0000 m? N? #9 >0 0:00 5000 000 Ex» «:0xoou pm 0 00.0000 00.000< 00.0000 00.0000< 0000-0-2 V? H? 09 N? Nana awed omv own 92 100 cm-1 to 1500 cm.1 from the exciting line. Generally, only the 01 (935 cm-l), 02 (460 cm-1) and v4 (626 cm-1) bands were observed, as the 03 (1110 cm'l) band is very weak in the Raman and was observed only at high concentrations (e.g., :_l fl_in water). These three bands appeared in all solvents studied. The only new band readily observed was m 470 cm”1 in acetonitrile, tetrahydrofuran and pyridine as shown in Figure 20 for 1.0 M_NaClO solutions for 4 the first two solvents. The splittings observed in THF and ACN in the 400-500 cm-1 region were further examined as a function of concentration and the data treated with a curve fitting computer routine. The equation employed to resolve the overlapping bands shown in Figure 20, is a Lorentzian-Gaussian product described by Irish, gt: El: (125): I = Io{exp[(-'J-Uo)2/2o21}{1 + (3-3012/021'1 where I is the arbitrary intensity at frequency 3; 3; is the position of the line center with maximum intensity, Io; and o is the variance, which gives the halfwidth when multiplied by the factor 1.46. This equation was employed to fit the data by using either a non-linear weighted least squares analysis, KINFIT (98) or a sequential simplex method, PEAKSBF (126). The results of this data treatment are listed in Table 12 and a sample of the computer output is shown in Figures 21 and 22. Presently, PEAKSBF is limited to resolving only two overlapping bands, whereas, KINFIT may resolve two, three or four overlapping bands. 93 TETRAHYDROFURAN ACETONITRILE INTENSITY [ I 1 I i 500 475 450 425 400 A FREQUENCY (CM'I) Figure 20. Raman Spectra of 1.0 H_Sodium Perchlorate Solutions in Tetrahydrofuran and Acetonitrile from 400-500 cm'1 94 H -000 0000.000 0.000. 00.000000000000000. :0 000.002.0000 050000 m 0.0 00 000.0002 0.00200. .00 0.000000 30¢ I ccc 7¢0=ucac>2¢ahwb 2— whdaCJICnuo !:hccv ac.- eeocoooooooetootoooooooo 07w Fm----m---'m----m--"W'---W"--V'--'m'---V'-'-v----U---'m--'-w-lv'-r'-'-U---IU--'-VII'l U U I I-"--'-I-"-— mu nocc . .muuuuw xuuooc . ccsnccc guarcc * uxncc . . Ctr x ywxuuu ccc xxcc urn x "vary :0 u u xx! . C" .C( x Ctr ~ m. "x . x run (0:20; 0 — Ox . u u >0 f 0 n w — Cx 0 0.. xu _ V x u — Cx C u — x 0 — Cx C _ 0 x x .0 U C u _ c 0 0 x x 0 0 c 0 0 0 0 V C X} H 0 x 0 0 0 0 c C 0 x 0 .. . 0 0 c 0 0 c 0 0 x . 0 V C x u 0 c 0 p xC — 0 — p C , xxulflfl — U x CCuCC n O u 0 . 0 xx x 0 0 x 0 ~ 0 m C — w - 0 w 0 ux c 0 0 x 0 0 0 _ V C u 0 u x 0 0 C u — — xC . 0 0 - 0 . 0 Nm---'U'-'-w-"'m--'-m--"W----w--'-m'---v----ui---v----m-'--U'--'v-'I'm'-'-T--'-m--"u'|"T"-'U--"h - > than: >3 x drama wzcv w:» 20 mm: 020cc cahagzuacc c:c 04020:020axu (a u2 2:. 020cc ou»«0:u0.noou~nn. u-ac» at» he m24<> .6000: cc ._.~.x m0 0qu00au> ocoucoo. u pzuxuaoz0.noou¢oe. n rte—a 01» >4 w:0¢>.ncoucce. u 000. u:» 04 u:.<> .0..oc:x.x ac 00.0.x v0 vvcocq I ~...I. u. u . up u u no .IIIIII . O . III II’.,.I I b t I D b 0 I p p n ..... IooIIIIIIII II II. III-90¢ . III’..I00 I I II. 3!. . Io IIIIIIIO. I III..I.!0.¢. . It...l..lcu I.....II..II. . . II .o.)!.to. 95 I I It!" I I .2396.- I I II. III—Ion I II II. ’0... I II II. 3000‘. .I II 93.33,. I I I II. ’2’. I I I II.IIII.III I I I I2; I I I I3};- I I I 0331’... I I I 3.113. O I I II.IIIIIII. ' I IT'IIIIII . I I223... I I I II.IIII¢II. I I I II.IIIIIII. I I I III-Iago. I O I OOIIIOI‘II O I I II. IIIIoII. .. 0.50.1... II II.§IIII. I I IIIIIIIII. I I II..IIIIII. I I II.III¢I¢I. I I 2.3.1:. I I IIIIIIIII. I I I II.IIII¢II. I I I II.IIIIIII. I I I II.IIIIQII. I I I II.IIIIIII. I I I IIIIIIIII. I I . . 0......03. I . I I 2......3. I I I 3.335.. I I 2.8.310. I I II.‘IIIII. I I 3.80.3... I I 2.3.3:. I I 3.333.. I I IIIIISII. II 3.33:... II II.IIII;I. II II.IIIIIII. I II.IIIIIII. I II..IIIIII. I I I II.III..II. I I . I II.IIIIIOI. I O I II. .IIIIIII. II I I333...- I I. I 3.3-3... I I I II.IIII~.I. C I I II.IIII.III I I I I II.IIIIIII. I I I 07330:. I I I ITIIIIII. I I I II.IIIIIIIo I I I 3.83.11 8 II II..I,III I II I130. I II II..IIIII. I I IICII‘II. I I III!:I. C I IIoiII II..IIIoOQo I I II..I-I.In. I I. .3083. I .- I00IIIOI I I.IIIIIIOQ II. ‘1‘. . 09800.08 III." OIaIIIolI‘ IooIIIIIQII II.IIIIIIII II.‘I¢IIQI II.‘I'II!I 3.18.0. III-III. II...IIIIO. II.IIIIIIII II.II.IIII. II.IIIIIII. II.IIIIIII. 0.0.0IIIIII II.IIIIIII. II.I.IIIIIO IILIIIIII. II. IIIIIII. QC. I..IIIII II. 30.1:- ‘0II..I'.OI II. IIIIIIII .II3IOFIOI IIIDIIOIOII II.:I0 III. II.IIIIIIII II. 87.... II.III.IIO. II..III.IIoo II.II~IIII0 OI. 'IIIIII 000:0I0flflo II.III.000. II. IIIIIII. I233...- II.II..III. II. IIIIIII. II.III.III. GI. IIIIIIII II.IIIIIIII II. IIIIIII. 00.50000! II.IIIIIcI. II. III. II. II. IIIIIII. II. 0.0:... II.II¢.9IC. II. III. I. II.I.I.II¢0 II.’I1¢I II. III-(I I. II. 0301‘ II. .101-Io IOIIII’IOI I III-III o. I I. III: to in ate 1Perchlor PEAKSBF Analysis of 1.0 M deium Tetrahydrofuran from 430-498 cm Figure 22. tal data I experimen x: ’ calculated'data O = eXperimental data; X = calculated data 0 ABOVE BELOW 96 Table 12. Computer Analysis of NaClO Solutions in Tetrahydrofuran 4 and Acetonitrile Using PEAKSBF and KINFIT Conc (M) Position (cm-1) Intensity 61/2(cm-1) 1470/1456 ACETONITRILE/PEAKSBF 1.00 459 213 31 ‘ 0.40 476 85.6 19 1.50 456 214 24 0.60 472 128 22 2.00 455 142 20 0.80 468 114 29 TETRAHYDROFURAN/PEAKSBF 1.00 453 307 22 0.75 470 231 22 O 80 453 272 20 O 83 470 226 22 0.60 454 157 19 0.74 471 116 19 ACETONITRILE/KINFIT 1.00 457 167 18 0.36 476 59.3 22 1.50 456 217 16 0.59 472 127 15 2.00 455 159 16 0.62 470 98.2 15 Table 12 (Continued) 1.0 0.80 0.60 453 470 453 470 453 470 97 TETRAHYDROFURAN/KINFIT 299 218 267 212 157 106 16 18 14 18 13 16 0.73 0.79 0.68 98 The data in Table 12 suggest that the m 456 tm'1 band is characteristic of free 0104- ion whereas the W 470 cm-1 band is assigned to ion paired C104 ion. These assignments are based On the fact that the ratio of intensities, 1470/1456’ increases with increasing concentration dramatically in acetonitrile and very subtly in tetrahydrofuran. Relating this to the 23Na NMR data, in acetonitrile, the chemical shift varies upfield m 1.5 ppm, indicative of dramatic changes in the chemical environment of the Na+ ion due to a cation-anion interaction. Hence, we would expect the Raman bands characteristic of free perchlorate ion and complexed perchlorate ion Q to also change dramatically. In acetonitrile, this ratio, 1470/1456’ varies from 0.40 to 0.80, reflecting an increase in the relative amount of cation—anion interactions. 0n the other hand, in tetrahydro- furan the chemical shift moves upfield only by W 0.15 ppm over the concentration range 0.60-1.00 M, These data imply very little change in the Na+ ion environment; equilibrium between free sodium ion and complex sodium ion is almost established. The static nature of the 1470/1456 ratio confirms this hypothesis. In addition, Figure 20 displays a new, weak, polarized band at m 412 cm'1 in the tetrahydrofuran solution. A corresponding band is not observed in the acetonitrile solution, perhaps due to a strong, broad solvent band at N 380 cm-1. 1 and 935 cm'1 bands appeared symmetrical, with no new bands present in the region 100-1500 cm-l. In both solvents, the 626 cm- Acetonitrile also affords the Opportunity to monitor the disposition of solvent molecules in solution as well. The C-C . . . . -1 symmetric stretching mode for acetonitrile occurs at 919 cm . Janz (80,81,127) has noted a new band at t 928 cm.1 for AgNO3 99 solutions in acetonitrile which he assigns to complexed solvent. The Raman spectrum of NaClO solutions in acetonitrile as a function 4 of salt concentration is shown in Table 13 and Figure 23. With increasing concentration, the ratio of complexed to free acetonitrile, I I919, increases as would be expected. The half- 927/ width (VI/2) of the 934 cm"1 perchlorate band seems to decrease with decreasing concentration which may be indicative of complexed perchlorate ion going to free ion. As mentioned earlier, both the 456 and 470 cm.1 bands appear in a 0.5 flNaClO4 solution in pyridine. From the 23Na NMR data, it was concluded that contact ion pairs are present in this solution; these Raman data lend support to that conclusion. More interesting was i the marked broadening of the 934 cm-1 perchlorate band in this solution, which may also be indicative of a cation-anion interaction. An inspection of Figure 19 suggests that in addition to splitting upon interaction, the new 460 cm.1 band(s) should also become infrared active provided the complexation interaction is strong enough to perturb the Td symmetry of the C104- ion. The far infrared region from 100 cm.1 to 600 cm.1 was scanned fbr NaClO4 solutions in ethanol, dimethylsulfoxide, propylene carbonate, dimethylformamide, pyridine, acetone, acetonitrile, tetrahydrofuran, tetramethylurea and tetramethylguanidine. No new bands were noted in any of these solvents except for the sodium ion solvation band at N 200 cm-1. Therefore, the vibrational studies suggest that the Na+-C104' interaction is not capable of producing strong changes in the C104- ion spectrum as did the covalent metal perchlorate salts 100 Table 13. KINFIT Analysis of Raman Spectrum of Sodium Perchlorate Solutions in Acetonitrile from 900-950-cm‘1 Conc (M) Position(cm-l) Intensity v (cm—1) I /I 1/2 927 919 920 92.6 8 2.00 927 116 16 1.25 934 209 10 919 147 7 1.50 927 99.2 18 0.67 934 150 7 919 213 7 1.00 927 110 18 0.52 934 153 S 101 H -Eo ommncom Eoum oflwuuwcouoo< cw mcowusHom oumuofigonom ssfipom mo mnuoomm cmemm .mm enamfim .729 55:85. a a...» new as see e5 93 _ _ _ - _ _ 01/. . _- \ M I. a j a 3 N Ah I. a .A a a flow 2.3 23 102 studied by Hathaway and Underhill. In these electrolyte solutions, only the 460 cm.1 Raman band is strongly affected by cation-anion interactions, while some line broadening is noted in the 935 cm-1 Raman band. These changes were quite apparent in NaClO4 solutions in acetonitrile, tetrahydrofuran and pyridine. However, a reexamination 1 of the 460 cm- Raman band in acetic acid, ethanol, water and fermic 1 acid shows marked asymmetry on the high wavenumber side which may be indicative of a weak band at m 470 cm-1. However, befbre these bands are studied in detail, we must be assured that the noted tailing is not an instrumental artifact. In addition to these vibrational studies, 35Cl NMR was employed to monitor the disposition of the C104 ion in solution. Complexation of the C104- ion should effect the quadrupolar relaxation mechanism of the 35C1 nucleus and result in a broadening of the resonance signal; in addition, such complexation may alter the electronic environment of the 35Cl nucleus and result in chemical shifts. Tables 14 and 15 diSplay the 35Cl linewidths and chemical shifts for NaClO4 solutions in water and several nonaqueous solvents. In all solvents, the 35C1 linewidths seem to be the same within experimental error (m 30 :_5 Hz). Similar solutions of LiClO4 exhibit marked 1/2 values ranging from 35 30 Hz to 170 Hz in a given solvent (128). The C1 chemical shifts linewidth changes with concentration, with 9 all seem to be identical within experimental error (N -1041 :_5 ppm) except for concentrated solutions in water. The relative insensitivity 35 of C1 linewidth and chemical shifts for sodium perchlorate solutions suggests the cation-anion interaction does not significantly perturb 103 m.“ mNoH- o.H mQHxOm43m4>EHmZHQ m.“ emofi- o.H fleaav emaem flag eceu zzam2Ho m.“ eeofi- o.H mzoemo< m.“ meoa- o.H neaav emagm flzv ecou qumHHZOHmu< m.“ emofi- o.~ aHu< qumoa m.“ Nmofi- o.~ 85.5% m.“ meofi- o.H m.“ mica- o.m m.” asofi- o.m magma peacm fizv ecou mmezemzHo muco>flom msooscmcoz :fi wcofipsaom oumpofinohom esfiwom pom Lppfl3ocflq do away nuwflzmcflq mm m.o m.“ me, o.H zmzemzHo .mH eaeee mm 105 the 35C1 nucleus. In contrast, marked changes in 35C1 NMR linewidths and C104' Raman and infrared bands have been observed for LiClO4 solutions in nonaqueous media (128,129,130). Hence, relative to the Li+-C104 interaction, the Na+-C104- interaction is significantly weaker as shown by 35C1 NMR, Raman and infrared measurements. Altogether, the data suggest that in NaClO4 solutions, weak cation-anion interactions exist; the 23Na NMR data strongly suggest that this cation-anion interaction is contact ion pairing. The problem is to rationalize contact ion pair formation in a high donor, high 1 dielectric constant solvent like water even at low salt concentrations 5 (~0.1 fl) and to explain the linear upfield shift in the 23Na resonance with increasing NaClO4 concentration in water, ethanol, methanol, and i fbrmic acid. This problem can be resolved by considering an interaction between the C104' anion and the solvent molecules or a lack of such interaction. The chemical shifts of solvent protons are generally altered by the addition of electrolytes. Among factors known to contribute to such changes are polarization of solvent molecules by the ions and modifications in the solvent structure produced by the ions. The polarization effect usually leads to deshielding of the solvent [arotons and hence a low field shift, unless magnetic anisotr0py caffects are present, which may cause a high field shift. The structural effect results in a low field shift if the solute is a net structure lulker, and a high field shift if it is a net structure breaker. Since the C104' ion is known to be a powerful structure breaker, we “mtlld expect a low field shift if the polarization effect dominates “K1 a high field shift if the structural effect dominates. 106 While studying the 1H resonance of the hydroxylic solvents water, methanol and ethanol, Krumgal'z, gt, a}; (131) noted high field shifts of the OH protons when NaClO4 was the solute. In another study, they found the protons in acetone and acetonitrile to shift to lower field with the addition of NaClO4 (132). In the former case, the data were rationalized in terms of strong structure breaking, whereas the polarization effect was dominant in the latter case. Coetzee and Sharpe (130) noted that the C104 ion did not effect the C-H stretching frequency of CHSCN at all, whereas other anions caused significant changes. They concluded that C10 ' does not interact with acetonitrile, 4 confirming the 1H NMR data of Krumgal'z. Hence, by invoking the concept of structure breaking, the upfield shift problem may be rationalized. The upfield nature of the shift 4 . + . cannot donate as much electron den51ty to the Na ion as could the itself is the result of a ctaion-anion complex where the C10 ion 4- complex may exist in water because the C104‘ will not be hindered by solvent . . . . + solvent. Despite the high dielectric constant, a Na -C10 structure about the cation in its quest for the Na+ ion. In this respect, it is interesting to note that the linear upfield shift only occurs in the hydroxylic, structured solvents water, methanol, ethanol and formic acid. These shifts may imply that when contact ion pairs are formed in these solvents, it is by collision. Also, if the energy of interaction is less than kT,the probability of their occurence would vary linearly with concentration. In contrast, the energy of the classical ion pair formed in acetonitrile, tetrahydrofuran, etc., would be greater than kT and result in a non-linear dependence 107 on concentration. The fact that only a slight upfield shift in the 23Na resonance is noted in tetramethylurea and dimethylformamide implies that contact ion pair formation is not favorable in these high donor solvents. CONCLUSION Although further investigations are in order, the use of spectro- scopic techniques to examine the upfield shift problem has at least displayed that cation-anion interactions are responsible for the observed shift. Moreover, speculations as to the types of species present in solution were advanced. This investigation emphasizes the idea that to fully understand the role of solvent in a chemical process, a sound knowledge of solute-solute, solute-solvent and solvent- solvent interactions in that solvent is required. APPENDICES APPENDIX I DESCRIPTION OF COMPUTER PROGRAM KINFIT AND SUBROUTINE EQN FOR THE CALCULATION OF ION PAIR FORMATION CONSTANTS BY THE NMR TECHNIQUE DESCRIPTION OF COMPUTER PROGRAM KINFIT AND SUBROUTINE EQN FOR THE CALCULATION OF ION PAIR FORMATION CONSTANTS BY THE NMR TECHNIQUE The numerical calculations for the ion pair formation constants were performed on the Control Data 6500 digital computer using the KINFIT program (98) by manipulating SUBROUTINE EQN. This appendix lists a typical deck for KINFIT analysis and describes the construction of this deck. Recall equation 16 which expresses the observed chemical shift, 6 , in terms of the total salt concentration, CM the chemical obs T’ shifts characteristic of free and complexed metal ion, 6F and 6C respectively, and the ion pair formation constant, K. -1 :_(1 + 4KC¥)1/2 M ] (6F - 6C) + 6C (16) ZKCT aobs = [ _ M dobs’ ”(1) - 6C’ T we obtain the FORTRAN equation for use in SUBROUTINE EQN for each By letting S U(2) = K, XX(1) = C and CONST(1) = 6 F data point. 8 = ((-1,0+SQRT(1.0+(4.0*U(2)*XX(1))))/(2o0*U(2)*XX(1))) *(CONST(l)-U(l))+U(l) . (28) 108 109 Note that we assume a constant value for 5F and that 6C and K are unknown. In order to fit the calculated shift (the right hand side of equation 27) to the observed shift, the program may vary the values of 6C and K. Hence, the number of unknowns, NOUNK, equals two as does the number of variables, NOVAR. The first data card contains the number of experimental points in columns l-5(FIS), the maximum number of iterations allowed in columns 10-15 (FIS), the number of constants in columns 36-40 (PIS) and the maximum value of (A parameter/parameter) for convergence to be assumed (0.0001 works well) in columns 41-50 (F10.6). The second data card contains any title the user desires. The third data card gives the value of CONST(1), SF, in columns 1-10 (F10.6) in ppm. The fourth data card contains the initial guesses for U(l), 6C, and U(2), K, in columns 1-10 and 11-20 (F10.6), respectively. The fifth through N data cards contain XX(1), concentration in columns l-10 (F10.6), the relative variance of XX(1) in columns 11-20 (F10.6), XX(2), the chemical shift at XX(1) in columns 21-30 (F10.6), the relative variance of XX(2) in columns 31-40 (F10.6) followed by the same parameters for the next data point. Hence, each card may contain two data points. The relative variance is error in each measurement normalized to each other (e.g., if the concentration error is :_0.04 M_and the shift error is :_0.1 ppm, the relative variance of the concentration is 0.04/0.04=1 and the relative variance of the shift is 0.1/0.04:25). If there are an odd number of data points, the last data card may contain that one point with columns 41-80 blank. If no further data is to be analyzed, the next 110 card after the last data point(s) should be a 6789 card. If more data set is to be analyzed, the next card after the last data point(s) is the number of points, number of iterations . . . etc. for the next data set. Generally the most common error in using this program is an error MODE44in an address of the SQRTE subroutine. Usually, this implies that the initial guess for K is quite inaccurate. Before using this program, the user should see reference (98) or the materials of CEM 883, Chemical Kinetics, to become familiar with the mode of operation and further application of program KINFIT. 33°C 111 ItpMfi P1665“ 12117) 1, ”nqu DOPOVcCN45000.1201.1c£J.n..q,. FTN. ATTACH1K1KF1T¢K1NFIT1 . LOAD(K1NFII) L60. 1 0000000000000003101900 snanourvvf ~09 suastolzen unsun-M5.uanuuaun§?z' FGNVOM VWII' .[Tu”v.1ThW"o1J“T IADQRINCP. N09T NOVAQQVOHNKQXoUoITNAKQFON ‘52; luTYoIrCT-IofVOQ‘KlqoI‘QoFOQQITYPOIIoUXTYPQH;lIofOPQFOQFUOPQZLOTnOEFQN Ztnvol_.xsr.7.1r.t.2.111.v.nv.vrct.anT.cn~.q nlurnqlow x14.1uu). ”(771. utx(4-1001. xx-Q #4 «a .g ) "Mllflfi D .71 > 1') point 9 FnouA [\(\ 53 no 7 to = I sa¢7u1.tn» CI‘NT {I 0“- CoHTImwF no 9 11 = Iota: 10! = In 0 l sA(xnl.rn) = .HI 4 satrnl.ln) o sAtInl.IQ) «'4 QafiJ—c—o—d—nb to a'vJE ~CODJDDJ \ Dfl ark", ‘00 FF. a"; pugwr in. r:. in 10 rnnuar l?V*.‘l 9th" [0%. lk‘. In? DO’KY 10“. ‘FF. "an 106 FOQVAT (81%| 7a = O. In? : [11 o i 102 = l (‘0 In [AI = l-II)n HH = 0.0 AM : (Ff - WI)/WV o l. as no 15 In = I¢¥.:nc.l H = Flflfif (II) no I] I“ 1 0. I”! ”(10‘ = m‘('”lol‘! l] COHTIHNF IF (Inn - /) av.¢'.l? ]? IF (1%“ - 7i Flo"i.l~ 11 no 10 h! 120 FORMAT (IN/o “Iowa ”an: [1 In THIQ pvtnfv) 119 5 AIB/A1419(H(4)°EXP(AI§/dl6)1/(10417/ , ’ 2 U I ‘ V: I. o l\ H o 0 \l \l / H I. 7 AU 0 I. \l 9. «I ‘- 1: II. uh. \l I. 0 9 on“ Q 0 ) 1| 9 D 9 A 1.. 1R nac. Q1. 1 I n a I. 0? 7. «I? :40. A. )0 o X o 9. a DA. Co. o co nos). 4 :1 A l. 1 o l a/ 9a Al 0 7% 00. c o! 1 A DA 63% $1. 0) .44 ? )? 41“.? 0 PA n .I. o 01971¢147.\rll ll, 9 9 Q- R 0 011.0 7 R Cc ell ll 11* Quill». 14‘ "A {wet I) II?) on.“ 0 ”U '19.. vlln '11 ll \l/\I 01/1 .0 llx o o a? O 9 11).”) A I 0 9n. OI! Cu (19 /(fiQ/F 1A RI. cl 1.! =1 ? I. In ( .l. 1| 0((1UI\(1) nl 1 H O :11) :Kllko C1n.‘ \. 2V.“ 2A.! 2‘ 2x0 .wofi.uuh) . Hl) V2 V4 23118 1 = S x U... .(U( _(1 . o o .9 13c s) a. nr 1 QKK‘O K .111 HUI. X 2 o 1 1:2 IOIO‘GH 0‘ .l E21410F 10 .qu/an .11?r|P.flfll-..U'I?CP. WU.” n n. C. ..rll M... FW:F. .J. WP .?(1.2H\(D\ ADMIN . .1”. (I (OH. 0'2 (OH. 1. o 7.......(fi.”h7?111~1011:1‘l:.fllUK .I. n U1 ”'11....” I. 0(AM,)QA M1V 1V1nvnv DAVIW. onnoasvolu GK 0 NH I. 1.1M. 1“ .l H 3: == = = == K 12.27....1 I.” .l 71...“ .10 011.1“ '10.... ======01Q§|111= .12 I?..l.|.:3W:I 3": =rln‘l\ TI: ._?l\==7l=2((::TIT=1I|I\(= u—TYI(= .1...P~«/l\n.1.vll.a :3H211H' «1.! fl ‘73h567n 9 NJ! 7.... V N N 21./4.56.5 7.. II. “1.1... 71 1V cl”? (3 N IIIIIIIIIIFHHunt‘Fprnnlfiin—PWInlwnrpr 6?.nnqin—rFFq110nFIIWMucuvnwwnl‘nFID 1000(“N0I0U0‘U nu AAAAAAAAAAVIHACII'YCKRDIRKCRDIIRKCF“0'31!“KCFIKIIIIIIO‘nleerl'lnv-vlcnxrclhvXclr. l a .I. 56 4730. 0 .1 11.567 2 345 0 .108 30,0 . .I- 2 3 l 0 11 all}. 2 2 2222 0 000 0 07“ 27.3 3 3 3 0 3 333 3 31 II. 3% 39 96 36 37 39 40 41 81 4? 41 4.1. as 97 46 (.7 310 99 49 '11 m1: V—L 3 120 i x rm . '2 .-. «unau- vnwan "21.-m. .Ls .. (~th - vuuan 9 0 falls 2 ['- COMT [Ml [F 101 = K1 DO 16 I“ = 19109 §°(lnloTC) = XX(I“) CGHYIHHF IF (lfi-h.) 17.17.53 H” = AM : O 1 F - WI)/MK o l. O J'. 1 ) 10041.“? “I END-dd I o OI .100 7(1w1) 3 3 4 D 3.~ x a Jxxcno DO 9 Jul ll _fia A—Hhflb" DZ-‘d ~ 179179h6 ~ II N". d «E O U 0- XI. —‘ 933‘ = 1.100 = HIIWQ) A kw r—e:<~ u , _ a—a. -u/ 7'1 ~ A-d "AA II-d ll Ilnbll H A-IIII-O‘D II ',~ V— “o’ 37OQQ.Q? (I, I”: : 90‘ Hl'flfil-QQJ ~3.a1.aa "R—vadaxufixD-filxx~fihmnD—ww—fl—un3-o ~Ofl3bfiflbJfiLflDfiflflSDDDODflDODDJ) a(7u5) hut" H [MHF 7“ . 10 HQ - 41) J?-07.&7 Kg - a) 16010.4“ 0. O 40 KM“) 11°o110.1w9 q = Kl = K? «on — «on . 1 Q1 : 94 7" - 0. an m an Ml =w1) vnw = 0 TT = 11 . . On Q) ["2 : 2.10M A316 = AqS(-1-’II-\I‘(I‘9H I" (mm - .rm ) shew?" '7" = K 7F (TY-“00.) 10049.“? rrw‘HH' no :9 v0? = 1.309 121 I polar R1. (0'. 64(1‘“"7‘ S3 FOP-”LT (l,:/.-§"1‘\I...'4C‘f,f.. -‘..‘;.T?: (>.L~1,).H) otwr/v : Ka(l~' ’1 £34 (2M1 hill? 033‘ T 3v» 55 FOWUQT I'M/o//0|¥.’V WNWEW'MzYkfi'H"*V rOQDDIMATE¢tho”CALCULATF3 1v (QODHYNQTF*-4!9*UFCIWHAI*9/’ Kfih = «ma 0 l H = WY 5‘! = I. (In T” ‘(0 sq nn(’(:\y:) : .11 r» = AH(AH) - AI ‘ no[»r g7. u.A”(P”\.\!.'§ 97 anunr «la/.?17.A.~<-Fl>.fiollYoFl’-6~l?x9El?-5’ IF (U - wJ) 89.6w.=0 SQ H = U o NK A“. = «N ’ ‘0 an 10 1a 50 pork} «0 60 Fn'H’AT‘l“!ol“(.*1*).?fl(§+¢.+[u)./) HRH : §J(1) ARF = «0(1) no 65 I = 2.‘w IF (Ann-AH(I)) 87.69.51 6’ ARO = AR([) 63 IF (ARF-AH(‘)) bqohboAh 64 59F = AH(I) Cnutxnur “\I 3 III! no 67 I = '0 4L 1” = (AQ(I)/(AHU-5WF\3000,o1,§ 1"! = (ANC(I)/(AHn-1\HE))uqq.o1_q D0 F” J = 1.10] TD!PT(J‘ = 1H 6H leTIMUF IF (IV-IN! 70.71.70 70 IDLCT(I”) = lfli . I°|07(IN) = 1'“) PRIhT 7?. AJ.IDLnT 72 FOQVAT (1H/~/. El?.6o 1x. 101a1) - no 10 71 71 ID[01(IH) = 1H: OQIKT 7 o AJ~I°I7T o wx 3 lax.%IG.P1(fl‘o¢+‘fi)./) . (aun-qur)/CQ. ou[\1 u 1. a“: 801 Fn)”AT“H/.Q‘Hr UAOHF HF THF Y CHOQUINATE AT THE ROTTOW = “061205, ought 204. D} 804 FnQVAT (IN/.vruF WiIH' 1F THV r CnnDUIMAIE AT THE Tap = «. 612.6! hVK : (AQW-AJ?)/0). DDVPT Q91. AV“ 321 pnovny (IN/.OTHF ugLHF 0? EACH IMCQFMFNT = “9E1?.6) (.0 = r A; = { I; ([51 - 2) QQ.0|.Q9 q? puHTr Q1 93 rnovnr (1H1.//.1y.ev Pvnnw'“1Tt*-‘Vo-“Y - CHQV‘ 1* ~7X-°Y - CUPVE 1p».7x.ev - FUQVF 1%./) an to q” 91 DJIKT 90 94 rnqwnT (1H1.//.}x.uv rnnanxnnTFo.Ax.ov - CHDVF 19.7x.6Y - CURVE 2* o/, Ian TO PO 9n Data? OR 9% FnDVATl(lHl-//ol<.*¥ CWnnWIUFTE*q6X-9Y - CHPVE 1’0/) 89 I“! = R! = 0. HI = n. RX 2 n. 122 n” F8 {:1 = 1.1;0.1)z IF? = {fl o IF." :: Vk'l 0 fi'; = -',\"-, .IL}‘.“ - AVA : 70(:(v.‘,/,. ‘4w9 5‘ = l”(I*l)-~‘-("" n1a)) IV (IFI - 1 ) ‘ ..I~ 1.34;. 183 Q! = a! G0 1'0 09 194 IF (Ir1-1-1nu; Hq.dk,fifi 49 0I : \I an 1n H) R". H’ (IF'l-l-Wu-‘T‘WH fun-17,43 R7 n“ = A! 82 (‘rWTIr-‘ut‘ 1F (JhQ-l) H*o719.llé RR 1‘ (Ian-9) 7fi1.7w;.7q1 701 DJYAT 704. n7. n! 70» FUUVAT (lH/.f12,~,g(,rlp.g) GU TO 70% 70? nwle 79%. av. 3:. n) 70:; FWDVQT '1"/0*-‘9.‘.‘~X.F1r’.fi.fii.fi|?.h) QJFfllzl7F3loq[#uv ADFA82ADF&20H’§.< AQEA3zbnfifii+fiK4uw Gn T0 70” 701 ODIN! 706. n/.n[. 9'. AK 70“ FORMAT (lH/G’fil)ohoi‘xor[aofififixo'El206Ofi‘QF-lao6) 70R IF (07—NJ) 707.703.7QQ 707 02 = 07 0 wK an TO 90 709 polnr an 07 : 1} 728 JflD 2 Jan . 1 GO TO R0 710 I“ = (AR(JAD)/lndn-nqc‘suqo..l.q V” = (HI/(ARD-M'C'))uo<)..1.t, IO = (RI/(AHH-ckryyouu..1.g ID = (PK/(AHn-AHF;)§QQ.,1.5 Dfl 711J= ‘Qljl - IDICY(J) = 1H 1 PONTIMHF 2 IF (IN-1n: 712.7)“,71: h IULCT(ID) = [Ht Y°IPT(IOJ = I”! Ioln1(1”) : 14: GO To 770 720 DQINT 79.07.13Lor an In 7?h 71% I? (IN-tn) 724.717.71; 717 YCLPT(ip) z gu. YDICT(IM) 2 !H1 19L"Y(lu) = 14: G0 70 77C 716 IF (I"-13) 714.714.71‘3 713 IDLCT(Y“) = 1H1 YOLPT(10) : [HY IDIPTtI") = 14: fin T0 790 719 IDlOT(IM) = lax VQICT(FO) = 1”! IOIPT(ID) : 14R I”LC1(VM) = 1%” N) IO 7 30 725 IF ()7-MJ) ~07.771.791 907 07 3.07 0 JK DUV’I 9?]. .‘.L( 9 DQIKT ‘9“). .‘.~"‘\]. -'."C‘I\f). r.u."\1 O? FqD‘JAT( '4/Q/lol,.\‘a)i-\;( I 3 Q. (.r 9 h. XoGL’ A v = o IQ'ZJ’K)FAX '11: “01Xorl?.fi) l l ' In E K a 'lx'EIZO‘NIO‘ Gn Tu Qqu 6] C’VITIMHF FHn APPENDIX IV DESCRIPTION OF COMPUTER PROGRAM SHEDLOV FOR EVALUATION OF CONDUCTANCE DATA DESCRIPTION OF COMPUTER PROGRAM SHEDLOV FOR EVALUATION OF CONDUCTANCE DATA Program SHEDLOV was used to evaluate conductance data using the CDC-6500 computer. Since this program is listed in the Ph.D. thesis of J. A. Caruso, this appendix lists and describes the construction of the data deck. The first data card contains a title of the user's choice in columns 1-56 (F7A8), a value for PER, which deletes those data points where the solvent conductance is greater than PER of the experimental conductance, in columns 62-69 (F8.2) and a value for CALCL in columns 70-72 (F13). If CALCL=0, the complete program /2 is used; if CALCL=1, then only Ao', C, A and C1 are calculated. The second data card contains ETA, the viscosity of the solvent in poise in columns l-lO (F10.0); DIELEC, the dielectric constant of the solvent in columns 11-20 (F10.0); TEMP, the absolute temperature of the experiment in columns 21-30 (F10.0); and ZERO, an initial guess at A0 in columns 31-40 (F10.0). The third data card designates control of the input data, IAM, in columns 1-5 (F15). For experimental data, IAM=O whereas for literature data, IAM=1. The number of data points, N, is in columns 6-10 (F15) and an integer which controls the extent of program execution, IAC, is in columns 11-15 (PIS). IAC=1 for weak acid and bases, whereas, IAC=O for electrolytes. 123 124 The fourth data card contains an integer which designates the method of experimentation, OPT, in columns l-S (F15). If solvent is added to a stock solution, OPT=1; OPT=O if stock solution is added to solvent. The fifth data card contains the density of the solvent, RHO, in columns 1-10 (F10.8); the conductance cell constant, KONST, in columns 11-20 (F10.8); the conductance of pure solvent, LSOLV, on columns 21-30 (FlO.8); the molecular weight of the solute, M2, in columns 31-40 (F10.8); the weight of original stock solution if OPT=l, WOSS, or the weight of original solvent if OPT=O, SOLV, in columns 41-50 (F10.8); the value of (grams of solute)/ (grams of stock solution), RATIOl, in columns 51-60 (F10.8); and the value of (grams of solvent/grams of stock solution), RATIOZ in columns 61-70 (F10.8). The sixth through N data cards contain for each data point the resistance of solution R(I), in columns l-lO (F10.5); the weight of added (not total) solvent if 0PT=1, ADSU(I) or the weight of added (not total) stock solution if 0PT=2, WTSS(I) in columns 11-20 (F10.5); and an integer signifying the end of a given set of data, IJ, in columns 21-22 (F12) on the last data point card of that given set. If more data are to be analyzed, the next card should be the title card of the next data set. When all set(s) of data are run, any integer, IM, in columns 57-61 (FIS) should appear on the last data point card of the final set along with IJ. 125 0.10 CONDUCTAKCF OF ROOIUM [GUIDE IN PYQIDINE .000160209.999859791 l S 7 l o h 7.. 9 R. O o 9 0 a 7 l R. . F a 6 0 1|. 0 O l 8 q 10 2 7.1.. 0 0 o 6 6,456 5 29217257796520 0 Q: COQQQCR,3449?H.O 3 9 QDROQAOR97OR§0 o 30.5300759918 o o .0 ? nanosecoooooalQo I. 00112222314579.1230 0 0 O LITERATURE C ITED (1) (2) (3) (4) (5) (6) (7) (8) (9) (10) (11) (12) (13) (14) (15) (16) LITERATURE CITED T. R. Griffiths and M. C. R. Symons, M01. Phys., g, 90 (1960). A. K. Covington and M. J. Tait, Electrochim. Acta, 12, 123 (1967). K. H. Wong, G. Konizer and J. Smid, J. Amer. Chem. Soc., 92, 666 (1970). S. Kopolow, Z. Machacek, V. Takaki and J. Smid, J. Macromol. Sci.—Chem., AZ, 1015 (1973). H. Wang and P. Hemmes, J. Amer. Chem. Soc., 2a, 5115 (1973). H. K. Frensdorff, J. Amer. Chem. Soc., fig, 600 (1971). B. W. Maxey and A. I. Popov, J. Amer. Chem. Soc., 90, 4470 (1968). E. Schaschel and M. C. Day, J. Amer. Chem. Soc., kg, 503 (1968). V. I. Chizhik, Strukt. Rol. Vody. Zhivm. Orgizme, k, 126 (1966). B. P. Fabrikand, S. S. Goldberg, R. Leifer and S. G. Angen, M01. Phys., Z, 425 (1963). J. N. Shoolery and B. J. Adler, J. Chem. Phys., 2;, 805 (1955). D. N. Nicholls and M. Swarc, J. Phys. Chem., 71, 2727 (1967). a) Z. Luz and S. Meiboom, J. Chem. Phys., 40, 1066 (1964). b) Ibid., an, 1058 (1964). A. Fratiello, R. E. Lee, D. P. Miller and V. M. Nishida, Mol. Phys., 12, 551 (1966). R. P. Taylor and I. D. Kuntz, Jr., J. Amer. Chem. Soc., 22, 4815 (1970). C. Deverell and R. E. Richards, Mol. Phys., 10, 551 (1966). 126 127 (17) C. Deverell and R. E. Richards, Mol. Phys., 16, 421 (1969). (18) H. Wennerstorm, B. Lindman and S. Forsen, J. Phys. Chem., 75, 2936 (1971). (19) C. H. Langford and T. R. Stengle, J. Amer. Chem. Soc., 91, 4014 (1969). (20) R. G. Bryant, J. Phys. Chem., 73, 1153 (1969). (21) G. E. Maciel, J. K. Hancock, L. F. Lafferty, P. A. Mueller and W. K. Musker, Inorg. Chem., g: 554 (1966). (22) J. W. Akitt and A. J. Downs, in "The Alkali Metals Symposium", The Chemical Society, London, 1967, p. 199. (23) E. T. Roach, P. R. Handy and A. I. Popov, Inorg; Nucl. Chem. Letters, 2» 359 (1973). (24) Y. M. Cahen, P. R. Handy, E. T. Roach and A. I. Popov,to be published. (25) J. D. Halliday, R. E. Richards, F. R. Sharp and R. R. Sharp, Proc. Roy,Soc. Lond. A, 313, 45 (1969). (26) A. Carrington, F. Dravnicks and M. C. R. Symons, Mol. Phys., 3, 174 (1960). (27) E. G. Bloor and R. G. Kidd, Can. J. Chem., 52, 3926 (1972). (28) o. Jardetzky and J. E. Wertz, J. Amer. Chem. Soc., 8g, 318 (1960). (29) G. A. Rechnitz and S. B. Zamochnick, J. Amer. Chem. Soc., 88' 2953 (1964). (30) M. Eisenstadt and H. L. Friedman, J. Chem. Phys., 44, 1407 (1966). (31) V. S. Griffiths and G. Socrates, J. Mol. Spectrosc., 27, 358 (1968). (32) R. E. Richards and B. A. Yorke, Mol. Phys., 6, 289 (1963). (33) E. G. Bloor and R. G. Kidd, Can. J. Chem., 46, 3425 (1968). (34) a) R. H. Erlich, E. T. Roach and A. I. Popov, J. Amer. Chem. §2£:’ 9%, 4989 (1970). b) R. H. Erlich and A. I. POpov, ibid., 23, S620 (1971). c) M. Herlem and A. I. Popov, ibid., 94, 1431 (1972). (35) (36) (37) (38) (39) (40) (41) (42) (43) (44) (45) (46) (47) (48) (49) (50) (51) (52) (53) (54) 128 G. J. Templeman and A. L. Van Geet, J. Amer. Chem. Soc., 94, 5578 (1972). A. L. Van Geet, J. Amer. Chem. Soc., 94, 5583 (1972). G. W. Canters, J. Amer. Chem. Soc., 94, 5230 (1972). F. W. Cope, J. Gen. Physiol., 50, 1353 (1967). C. A. Rotunno, V. Kowalewski and M. Cereijido, Biochim. Biophys. Acta, 11%, 170 (1967). D. H. Haynes, B. C. Pressman and A. Kowalsky, Biochemistry, (Q, 852 (1971). T. L. James and J. H. Noggle, J. Amer. Chem. Soc., 21, 3429 (1969). T. L. James and J. H. Noggle, Proc. Nat. Acad. Sci. (U.S.A.) 8%, 644 (1969). F. Ostroy, T. L. James, J. H. Noggle and L. E. Hokin, Fed. Proc., 31, 1207 (1972). T. L. James and J. H. Noggle, Anal. Biochem., 49, 2081 (1972). T. L. James and J. H. Noggle, Bioinorg. Chem., 2, 69 (1972). J. Andrasko and S. Forsen, Biochem. Biophys. Res. Commun., 52, 233 (1971). J. M. Ceraso and J. L. Dye, J. Amer. Chem. Soc., 95, 4432 (1973). E. Shchori, J. Jagur-Gradzinski, Z. Luz and M. Shporer, J. Amer. Chem. Soc., 93, 7133 (1971). J. C. Evans and G.Y—S. Lo, J. Phys. Chem., 69, 3223 (1965). W. F. Edgell, A. T. Watts, J. Lyford, IV and W. M. Risen, Jr., J. Amer. Chem. Soc., 88, 1815 (1966). W. F. Edgell, A. Lyford, IV, R. Wright, W. Risen, Jr. and A. Watts, J. Amer. Chem. Soc., 92, 2240 (1970). B. W. Maxey and A. I. Popov, J. Amer. Chem. Soc., 89, 2230 (1967). B. W. Maxey and A. I. Popov, J. Amer. Chem. Soc., 91, 20 (1969). J. L. Wuepper and A. I. Popov, J. Amer. Chem. Soc., 21, 4253 (1969). (55) (56) (57) (58) (59) (60) (61) (62) (63) (64) (65) (66) (67) (63) (69) (70) (71) (72) 129 J. L. Wuepper and A. I. Popov, J. Amer. Chem. Soc., 9%, 1493 (1970). M. K. Wong, W. J. McKinney and A. I. Popov, J. Phys. Chem., ZR, 56 (1971). M. K. Wong and A. I. Popov, J. Inorg. Nucl. Chem., 33, 1203 (1970). W. J. McKinney and A. I. Popov, J. Phys. Chem., 74, 535 (1970). P. R. Handy and A. I. Popov, Spectrochim. Acta, 28A, 1545 (1972). W M. S. Greenberg, D. M. Wied and A. I. Popov, Spectrochim. Acta, 29%, 1927 (1973). A. T. Tsatsas and W. M. Risen, Jr., J. Amer. Chem. Soc., 9%, 1789 (1970). J. A. Olander and M. C. Day, J. Amer. Chem. Soc., 93, 3584 (1971). E. G. Hohn, J. A. Olander and M. C. Day, J. Phys. Chem., 73, 3880 (1969). W. F. Edgell, J. Lyford, IV, A.Barbetta and C. I. Jose, J. Amer. Chem. Soc., 93, 6403 (1971). W. F. Edgell and J. Lyford, IV, J. Amer. Chem. Soc., 23, 6407 (1971). W. R. Robinson, A. Barbetta and W. F. Edgell, to be published. A. T. Tsatsas and W. M. Risen, Jr., Chem. Phys. Lett., 7, 354 (1970). A. T. Tsatsas, R. W. Stearns and W. M. Risen, Jr., J. Amer. Chem. Soc., 94, 5247 (1972). B. W. Maxey, Ph.D. Thesis, Michigan State University, East Lansing, Michigan, 1968. J. L. Wuepper, Ph.D. Thesis, Michigan State University, East Lansing, Michigan, 1969. P. R. Handy, Ph.D. Thesis, Michigan State University, East Lansing, Michigan, 1972. M. K. Wong, Ph.D. Thesis, Michigan State University, East Lansing, Michigan, 1971. 130 (73) R. E. Hester and R. A. Plane, J. Chem. Phys., 38, 249 (1963). (74) R. E. Hester and R. A. Plane, Inorg. Chem., 3, 768 (1964). (75) R. E. Hester and R. A. Plane, Inorg. Chem., 3, 769 (1964). (76) B. M. Gatehouse, S. E. Livingstone and R. S. Nyholm, J. Chem. Soc., 4222 (1957). "“““ (77) J. I. Bullock and F. W. Parrett, Chem. Commun., 157 (1969). (78) A. B. P. Lever, E. Mantovani and B. S. Ramaswamy, Can. J. Chem., 49, 1957 (1971). (79) D. E. Irish, A. R. Davis and R. A. Plane, J. Chem. Phys., 50, 2262 (1969). - (80) G. J. Janz, K. Balasubrahmanyam and B. G. Oliver, J. Chem. Phys., 51, 5723 (1969). (81) K. Balasubrahmanyam and G. J. Janz, J. Amer. Chem. Soc., 22, 4189 (1970). (82) G. J. Janz, M. J. Tait and J. Meier, J. Phys. Chem., 71, 963 (1967). (83) C. C. Addison, D. W. Amos and D. Sutton, J. Chem. Soc. (A), 2285 (1968). (84) D. J. Gardiner, R. E. Hester_and W. E. L. Grossman, J. Chem. Phys., 59, 175 (1973). (85) J. H. Roberts, A. T. Lemley and J. J. Lagowski, Spectr. Lett., 5, 27 (1972). (86) K. R. Plowman and J. J. Lagowski, J. Phys. Chem., 18, 143 (1974). (87) E. B. Baker, L. W. Burd and G. N. Root, Rev. Sci. Inst., 36, 1495 (1965). (88) 0. H. Live and s. I. Chan, Anal. Chem., 42, 791 (1970). (89) M. J. 0. Low, J. Chem. Ed., 47, A163 (1970). (90) H. B. Thompson and M. T. Rogers, Rev. Sci. Instrum., 27, 1079 (1956). (91) H. M. Daggett, E. J. Bair and C. A. Kraus, J. Amer. Chem. Soc., 75, 799 (1951). (92) (93) (94) (95) (96) (97) (98) (99) (100) (101) (102) (103) (104) (105) (106) (107) (108) (109) 131 T. Shedlovsky, J. Franklin Inst., 22%, 739 (1938). D. A. Burgess and C. A. Kraus, J. Amer. Chem. Soc., 70, 706 (1948). D. N. Bhattacharyya, C. L. Lee, J. Smid and M. Szwarc, J. Phys. Chem., 69, 608 (1965). C. Carjaval, K. J. Tolle, J. Smid and M. Szwarc, J. Amer. Chem. Soc., 87, 5548 (1965). a) V. Gutmann and E. Wychera, Inorg. Nucl. Chem. Lett., 2, 257 (1966). b) V. Gutmann, "Coordination Chemistry in Nonaqueous Solvents", Springer—Verlag, Vienna, 1968. U. Mayer and V. Gutmann, Struct. Bonding (Berlin), 12, 113 (1972). J. L. Dye and v. A. Nicely, J. Chem. Ed., $8, 443 (1971). L. S. Frankel, T. R. Stengle and C. H. Langford, Chem. Commun., 393 (1965). Ibid., J. Phys. Chem., 74, 1376 (1970). R. L. Bodner, M. S. Greenberg and A. I. Popov, Spectr. Lett., 5, 489 (1972). J. J. Lindberg, J. Kenttamaa and A. Nissema, Suomen Kemistilehti, 34%, 98 (1961). R. Figueroa, E. Roig and H. Szmant, Spectrochim. Acta, 22, 587 (1966). A. Luttringhaus and H. W. Dirksen, Angew. Chem. Intern. Ed. Engl., 3, 260 (1964). C. Beguin and H. Gunthard, Helv. Chim. Acta, 42, 2262 (1959). A. K. Covington, T. H. Lilley, K. E. Neuman and G. A. Porthouse, J. C. S. Faraday I, 69, 963 (1973). A. K. Covington, K. E. Neuman and T. H. Lilley, J. C. S. Faraday I, 69, 973 (1973). L. S. Frankel, C. H. Langford and T. R. Stengle, J. Phys. Chem., 74, 1376 (1970). P. C. Lauterbur, Ann. N. Y. Acad. Sci., ZQ’ 841 (1958). (110) (111) (112) (113) (114) (115) (116) (117) (118) (119) (120) (121) (122) (123) (124) (125) (126) (127) (128) (129) (130) 132 W. M. Litchman and D. M. Grant, J. Amer. Chem. Soc., 29, 1400 (1968). R. H. Erlich, Ph.D. Thesis, Michigan State University, East Lansing, Michigan, 1971. J B. Kinsinger, M. M. Tannahill, M. S. Greenberg and A. I. Popov, J. Phys. Chem., 17, 2444 (1973). A. D'Aprano, J. Phys. Chem., ZS, 3290 (1971). A. D'Aprano, J. Phys. Chem., 76, 2920 (1972). B. F. Prine and J. E. Prue, Trans. Faraday Soc., 6%, 1257 (1966). S. Minc and L. Werblan, Roczniki Chemii, 4Q, 1537 (1966). R. L. Kay, B. J. Hales and G. P. Cunningham, J. Phys. Chem., 71, 3925 (1967). P. G. Sears, G. L. Lester and L. R. Dawson, J. Phys. Chem., 60, 1433 (1956). D. P. Ames and P. G. Sears, J. Phys. Chem., 59, 16 (1955). B. J. Barker and J. A. Caruso, J. Amer. Chem. Soc., 93, 1341 (1971). S. Mine and L. Werblan, Electrochimica Acta, Z, 257 (1962). J. E. Prue and P. J. Sherrington, Trans. Faraday Soc., 57, 1795 (1961). R. M. Fuoss, J. Amer. Chem. Soc., 81, 2659 (1969). K. L. Hsia and R. M. Fuoss, Proc. Nat. Acad. Sci., U.S., 57, 1550 (1967). A. R. Davis, D. E. Irish, R. B. Roden and A. J. Weerheim, Appl. Spec., 26, 384 (1972). S. N. Deming and S. L. Morgan, Anal. Chem., 45, 278A (1973). G. J. Janz, J. Electroanal. Chem., 29, 107 (1971). R. Baum, private communication. Y. Cahen and A. I. Popov, to be published. J. F. Coetzee and W. R. Sharpe, J. Soln. Chem., 1, 77 (1972). (131) (132) (133) (134) 133 B. S. Krumgarz, K. P. Mishchenko, D. G. Traber and Y. P. Tseretell, J. Structural Chem., USSR (Eng. Trans.), 13, 373 (1972). B. S. KrumgaPz, Y. I. Gerzhberg, gt: 31:, Zh. Fiz. Khim., 44, 227 (1970). J. A. Caruso, Ph.D. Thesis, Michigan State University, East Lansing, Michigan, 1967. B. J. Hathaway and A. E. Underhill, J. Chem. Soc., 3091 (1961). WWITIWIWW1111111111111111mm!“ 3 1293 030617793