PART I: Synthesis and Characterization of Alkalides and Electrides that Contain l5-Crown-5. PART II: Alkali Metal NMR Studies of Alkalides. By Mary L. Tinkha- A DISSERTATION Subaitted to Michigan State University in partial fulfill-ent of the require-ents for the degree of DOCTOR OF PHILOSOPHY Depart-eat of Cheaistry 1985 ABSTRACT PART I: Synthesis and Characterization of Alkalides and Blectrides that Contain 15—Crown—5. PART II: Alkali Metal NMR Studies of Alkalides. By Mary L. Tinkham Ten crystalline compounds that contain alkali metal ions and l5-crown-5 (1505) were synthesized and analyzed. These compounds were precipitated from solutions of alkali metals and lS-crown-S in mixtures of dimethyl ether (MeZO) and either diethyl ether or trimethylamine. Analyses were in good agreement with the expected stoichiometries of K+(l505)2-Na-, n+(1505)2-Rb', Rb+(15c5)2-e', nb+(15c5)2-Na‘, ab+(15c5)2.ab' and Cs+(15c5)2-Na’ but deviated somewhat from K+(1505)2-e—, K+(1505)2-K-, Cs+(l5C5)2-K- and Cs+(15C5)2-Rb—. Pressed powder D.C. conductivity measurements indicated that these compounds are extrinsic semiconductors with apparent band gaps of 0.5 to 1.7 eV. EPR spectra of K+(1505)2-e- and Rb+(l5C5)2-e— showed a narrow line at g = 2.003 for each electride. Magnetic susceptibilities showed that a high fraction of electrons are unpaired in each com- pound. A ”jump" in the electronic xm of Rb+(l5C5)2-e_ between 120 and 140 K may indicate a phase transition. Tinkham, Mary L. Solid state alkali metal NMR was used to identify alka- lides that were synthesized both by the author and by others. Magic angle spinning (MAS) spectra were obtained for 23Na, 39K, 87Rb and 133Cs NMR. Eleven compounds were identified as sodides and each had the 23Na— NMR peak of in the range of -56 to -63 ppm. The NMR signal of the complexed cation was observed for the homonuclear sodides Na+c222-Na' and Na+(lzc4)2-Na’. 393 MAS-NMR studies of potassides provided the first NMR observation of K- chemi- cal shifts at -105 to -115 ppm. Quadrupolar broadening prevented the observation of a complexed cation signal both for K+ and for Rb+. The Rb- anion had a frequency dependent NMR peak at -187 to ~199 ppm. 1330s MAS-NMR was used to identify three alkalides and each had a 133Cs NMR peak at +24 to +29 ppm which corresponded to Cs+(1505)2. Some samples had additional peaks which corresponded to Cs+(1505)2-e- at +505 ppm and Cs— at -263 ppm. 39K and 87Rb solution NMR spectra of alkalides in MeZO are also reported. ACKNOWLEDGMENTS The author wishes to express her sincere gratitude to Professor James L. Dye for his guidance, encouragement and support throughout this study. I would like to thank the members of Dr. Dye’s research group, both past and present, for their cooperation, suggestions and aid. In particular, I wish to acknowledge Zexia Barnes, Steven Dawes, Margaret Faber, Odette Fussé, Rui Huang, Stephan Jaenicke and Iraj Behbahani for their help and moral support. Thanks are also extended to the newer members of the group, Lauren Hill, Jineun Kim, Mark Kuchenmeister and Joseph Skowyra. I also wish to thank Dr. Dye and his group for the fishing trips, the great Asilomar adventure and other group outings which helped to make working together much more enjoyable. Thanks go to the glass blowers Scott Bancroft, Manfred Langer and Keki Mistry for their speedy and excellent ser- vice. The secretaries, Naomi Hack, Carol link and especially Margy Lynch, are also acknowledged for their help. The cheerfulness and the helpful attitude of Dorothy Boetteger is also greatly appreciated. I also wish to thank the members of the electronics and machine shops. -ii- Special thanks go to Kermit Johnson for his help with the solid state NMR work at Michigan State University. Drs. Klaas Hallenga and Long Dinh Le are also gratefully acknowledged. A special "thank you" goes to Dr. Pat Smith at Dow Chemical Corporation at Midland, Michigan for his initial aid and continuing help in magic angle spinning NMR studies. Thanks go to Professor Eric Oldfield and his group at the University of Illinois for their aid in the 87Rb high field NMR work. Financial support is gratefully acknowledged from the Department of Chemistry, Michigan State University and from the National Science Foundation (grants DMR 79-21979 and DMR 84-14154). I would like to thank my family, in particular, my sister Linda, for their faith and encouragement throughout this study. Finally, I would like to thank my friends, Ruby Freed, Bob Kean and especially N.R. Nirmala, for their patience, understanding and support; for without their friendships, this road would have been impossible to travel. -iii- PREFACE The work covered by this dissertation can be divided into two areas -- each heavily dependent on the other. The first is the synthesis and characterization of new in- organic salts that have either an alkali metal anion (alka- lide) or a trapped electron (electride) as the negative species. In both cases the positive ion is an alkali metal cation complexed by macrocyclic organic ligands. Species identification is essential in the characterization of alkalides and electrides to correctly distinguish between pure compounds, mixtures and "junk". Historically, alka- lides and electrides have been identified via stoichiometry and optical spectroscopy. Although these techniques have proved satisfactory for most cases, some ambiguity arose for others. The second portion of this work describes the applica- tion of alkali metal NMR as an identification tool for alkalides and electrides. This nuclear specific technique was applied to all alkali metals except lithium. For the most part, it was restricted to the solid state where the "magic angle" sample spinning technique was employed, although some solution experiments were also performed. -iv— Because of the major differences in the background required to understand these two facets of this research, the material is presented in two separate sections, Parts I and II. Each section has historical, experimental and results chapters. A final chapter is used to relate the conclusions that can be drawn from these two sections and to present suggestions for further work. -v- List of List of Part I: Chapter A. B. 0. Chapter O’HWU Chapter HNUOW> TABLE OF CONTENTS Figures. Tables. Synthesis and Characterization of Alkalides and Electrides that contain 15- -crown— 5. One. Introduction. . . . . . . . Metal Ammonia Solutions. . . . Alkali Metals in Amines and Ethers. . . Properties of Alkalides and Electrides. Two: Experimental. Reagent Preparation. 1. Metals. 2. lS-crown- 5. 3. Solvents. . a. Dimethylether and trimethyl- amine. . . . . . . . b. Diethylether. Synthesis Techniques. 1. Glassware preparation. 2. Improved synthesis technique. Analysis. . . . . . . . 1. Hydrogen evolution. 2. pH titration. 3. Flame emission. 4. Proton NMR. Optical Spectra. . Pressed Powder Conductivity. Electron Paramagnetic Resonance. Magnetic Susceptibility. Three: Results and Discussion. Synthesis. Analysis. . . . . . Optical Spectra. . . . . Pressed Powder Conductivity. Electron Paramagnetic Resonance. Magnetic Susceptibility. -vi- PAGE .viii xi . 30 .30 30 30 32 .32 32 . 32 .33 33 38 40 42 43 . 43 .44 .47 .48 .48 . 50 .51 . 55 .60 .73 .82 .87 PART II: Chapter UGU> Chapter DO Summary List of Results from other Sources. 1. Differential Scanning Calorimetry. 2. X- ray Diffraction. . . . . . . 3. EXAFS and XANES. Alkali Metal NMR Studies of Alkalides. One: Introduction. Nuclear Properties of Alkali Metals. NMR of Alkali Metal Anions in Solution. Solid State NMR. Identification of a Ceside and an Elec- tride by 13303 MAS NMR. Two: Experimental. . . . . Sample Preparation. . . . . . . . . . . 1. Solution Samples. 2. Model Salts. . 3. Solid State Samples. Instrumentation. 1. At Michigan State University. 2. At Dow Chemical Corporation. 3. At the NSF Regional NMR Center. Three: Results and Discussion. 23Na NMR Spectra. 39K NMR Spectra. . 1. Solution studies. 2. MAS studies. a'7Rb NMR Spectra. 1. Solution studies. 2. MAS studies. 13303 NMR Spectra. and Suggestions for Future Work. References. —vii- PAGE . 94 .94 .95 .97 102 .102 103 .105 108 .115 .120 .120 .120 121 121 121 .122 123 .124 .125 .125 132 .133 135 .140 .141 .143 152 157 .164 FIGURE 1 2 10 ll 12 13 LIST OF FIGURES Models for Species of Stoichiometry M‘. Optical Spectra of Alkali Metals in Ethylenediamine. . . . . . . . Cryptand-2,2,2 and 18-crown-6. Log K versus Cation Radius in Methanol at 25'C. Packing of Na’C222 and Na” (solid circles) in the crystalline Solid Na’0222-Na'. . . . Born-Haber Cycle for Formation of M‘L-N‘. Vacuum Distillation Apparatus for 15-Crown-5 Purification. Synthesis Apparatus: H Cell. Analysis Scheme for Alkalides (A) and Electrides (B). . . . . . . . . Hydrogen Evolution Apparatus. Optical Absorption Apparatus. Optical Absorption Spectra of Thin Dry Films Made From Solutions of K’(15C5)2-K‘ (A) and K‘(15C5)2-e’ (B and C) in Methylamine. . . Optical Absorption Spectrum of 8 Thin Dry Film Made From a Solution of K(15C5)1 3 in Methylamine. «viii? PAGE .11 .14 .16 .19 20 .31 34 39 41 45 62 .63 FIGURE PAGE 14 Optical Absorption Spectra of Thin Dry Films Made From Solutions of Rb’(l505)2-Rb (A) and Rb’(l5C5)2-e (B)in Methylamine. . . . . . . . . . .65 15 Optical Absorption Spectra of Thin Dry Sodide Films Made From Solutions of Cs'(1505)2-Na' (A), Rb’(15C5)2-Na‘ (B), and K’(15C5)2-Na‘ (C) in Methylamine. . . . . . 66 16 Optical Absorption Spectrum of a Thin Dry Film Made From a Solution of Crystals with Stoichiometry KRb(1505)2 in Methylamine. . . . . . . . . . . .69 17 Optical Absorption Spectra of Thin Dry Films Made From a Solution of Cs’(15C5)2-K‘ in Methylamine. . . . . . . . . . 72 18 -Ln (current) versus 1/T for Rb’(1565)2-e'. . . 79 19 EPR Spectrum of Rb*(1505)2-e' at -103°C. . . . .84 20 EPR Spectrum of K’(15C5)2-e' at -97'C. . . . . .85 21 l/xm versus Temperature for Rb‘(15C5)2-e'. . . .90 22 l/xm versus Temperature for K‘(1505)2-e‘. . . . 92 23 DSC Trace for Rb*(15C5)2-Na'. . . . . . . . . . 96 24 K-edge X-ray Absorption Spectrum of Rb’(1505)2-Na'. . . . . . . . . . . . . . . . . 98 25 Diagram Illustrating the Motion of a Typical Internuclear Vector, ri., when a Solid is Rotated with an Angulag Velocity, 0r , about an Axis Inclined at Angle Br to Ho . . . . . . . . . . . . . . . . . . . . 111 26 133Cs MAS-NMR Spectra of the Two Crystal- line Compounds, Cs’(18C6)2-Cs’ and Cs*(18C6)2-e' at 44.24 MHz. . . . . . . . . . .118 27 23Na NMR Spectra of Crystalline Na’0222-Na‘ at 52.94 MHz. . . . . . . . . . . .127 28 23Na MAS- NMR Spectrum of Na*(1204)2-Na at 47. 61 MHz. . . . . . . . . . . . . . . . 129 FIGURE PAGE 29 39K NMR Spectrum of K’(15C5)2-K in Dimethylether (0.06 M) at 220 K and 8. 40 MHz. . . . . . . . . . . . . . . . . . .134 30 39K NMR Spectra of Polycrystalline K*(15C5)2-K' at 180 K. . . . . . . . . . . . . 136 31 87Rb NMR Spectrum of Rb’(1505)2-Rb' in Dimethylether (0.068 M) at 235 K. . . . . . . .142 32 37Rb NMR Spectra of Polycrystalline Rb’(15C5)2-Rb‘ at 65.4 MHz. . . . . . . . . . .150 33 133Cs MAS- -NMR Spectrum of an Impure Sample of Cs’(15C5)2-K‘. . . . . . . . . . . . 153 34 133Cs MAS- -NMR Spectrum of Cs*(1505)2-K‘ - at 23. 62 MHz. . . . . . . . . . . . . . . . .155 TABLE II III IV VI VII VIII IX XI XII XIII LIST OF TABLES Alkali Metal Solubilities in Various Solvents. Crystallization Conditions and Relative Stabilities of 15-Crown-5 Alkalides and Electrides. . . . . . . . . . . . Analysis Results of Alkalides and Electrides. . . . . . . . . . . . Results of Optical Absorption Studies of Thin Dry Films. . . . . . . . . Results of Pressed Powder DC Conductivity Measurements. . . . . . . . . . . Relative Areas of Rubidium K—Edge Absorption Threshold Resonances. Nuclear Properties of Alkali Metals. Calculated Chemical Shifts for Gaseous Alkali Metal Atoms and Anions and Observed Chemical Shift of M‘ in Solution. Results of 23Na MAS-NMR Studies. Results of 39K MAS-NMR Studies. Results of 87Rb MAS-NMR Studies at Three Frequencies. . . . . . . . Results of 87Rb NMR With and Without Proton Decoupling. Results of 133Cs MAS-NMR Studies. -xi- PAGE 10 54 57 68 77 .99 104 .109 128 .138 144 .149 .156 PART I: Synthesis and Characterization of Alkalides and Electrides that Contain l5-Crown-5. PART I CHAPTER 1 -- INTRODUCTION The synthesis of crystalline alkalides and electrides is a direct outgrowth of attempts by Dye and coworkers to extend the study of alkali metal-ammonia solutions to other solvents. This chapter will cover a brief history of the development of alkalide and electride chemistry starting with metals in liquid ammonia. The use of macrocyclic complexing agents to increase the solvent range will be discussed as well as the effect of these complexants on general properties. Finally, the synthesis and charac— terization of alkalides and electrides, prepared as powders or thin films, as well as precipitated crystals, will be reviewed. A. Metal Ammonia Solutions The unpublished notebooks of Sir Humphrey Davy show that alkali metal-ammonia solutions were prepared as early as 1808 [1]. These solutions were first described in the open literature by Weyl in 1864 [2]. Since that time, many papers have been written on the properties of metal-ammonia -1- -2- solutions. The proceedings of six international conferences [3-8] and a number of review articles [9-20] have been published. Although metal-ammonia solutions have been studied rather intensely, several questions remain unanswered. The area of metal-ammonia chemistry is still a very active field both on the fundamental and applied level. For example, the high reducing power of such solu- tions has been utilized to create new organic and organo- metallic intermediates [21]. Discussions of metal-ammonia solutions are usually restricted to one of three concentration ranges: dilute (infinite dilution to ~l.0 mole percent metal (MPM)), intermediate (1+8 MPM) and concentrated (>8 MPM). These distinctions are drawn on an arbitrary basis since the properties of the solutions change gradually with concen- tration. In general, dilute solutions have electrolyte behavior and concentrated solutions are metallic. There is a gradual transition from electrolyte to non-metal to metal as the metal concentration is increased. These transitions are usually seen in the intermediate range since the non- metal to metal transition is usually complete at a concen- tration of 8 mole percent metal. Liquid ammonia dissolves the following metals: Li, Na, K, Rb, Cs, Ca, Sr, Ba, Yb and Eu. It is universally agreed that the major species in dilute solutions are solvated electrons and metal cations. Indeed, one may avoid most of the cationic effects by introducing electrons by means of -3- ionizing radiation from an accelerator. The majority of metal ammonia solution studies involve alkali metals. This review will be restricted to such studies. Metal-ammonia solutions may be studied over a wide range of temperatures and concentrations. Even at the normal boiling point of NH3, -33'C, very concentrated solu- tions of the alkali metals can be made. Solubilities range from 15 MPM for potassium to 65 MPM for cesium for saturated solutions at -33'C. The saturation concentra- tions change very little with temperature, with the exception of that of (cesium, which is found to be completely miscible with NH3 near 25'C and above. There are miscibility gaps in the phase diagram for the other alkali metals. It has been speculated that all alkali metals would be completely miscible with liquid ammonia at the melting point of the metal. However, most metal- ammonia work is carried out at temperatures at or below the normal boiling point of ammonia. Even with this as a high temperature limit, the temperature range available is quite large. The eutectic temperature of lithium in NH3 is -184.6'C -- a freezing point depression of over 100 degrees! The concentration is surprisingly high at the eutectic temperature (20 MPM). The electrical conductivity of a saturated solution of lithium in liquid ammonia exceeds that of liquid mercury at room temperature. The other alkali metals in NH3 have con- ductivities almost as high. In concentrated solutions, the -4- electrical conductivity increases approximately as the cube of the metal content. The conductivity also increases in a nearly linear fashion with temperature, except for the most concentrated solutions of lithium and cesium. In these solutions a slight decrease is observed as the temperature increases. A second similarity of concentrated metal ammonia solu- tions and metals occurs in the plasma absorption edge observed by optical reflectance studies. This plasma edge occurs at approximately 10,000 cm.1 at the metal-nonmetal transition but shifts to the blue as the concentration is increased. The color (by reflectance) of concentrated metal solutions is a polished bronze. Magnetic data for M-NH3 solutions give rather confusing results. The static magnetic susceptibility shows a weak temperature dependence even in the most concentrated solu- tions. The derivatives of the EPR signal show an asymmetry of the low field maximum compared to the high field mini- mum. The change in the A/B ratio is reminiscent of metallic systems. A and B are the respective amplitudes of the low and high field lobes of the first derivative EPR spectra. These changes are observed at relatively low con- centrations. The magnetic behavior of metal-ammonia solu- tions, at any concentration, has not been explained in a satisfactory manner, although Thompson has put forth several models [19]. -5- Dilute solutions (concentrations (1 MPM) have a charac- teristic blue color at concentrations of 10.6 molar or more. The optical maximum for the solvated electron lies in the near infrared region at 6900 cm—1, but has an asym- metric high energy tail into the visible. This line is completely unaffected by the type of solute at low concen- trations. The absorption band is also observed when elec- trons are produced by ionizing radiation or when ND3 is substituted for NH3. The spectrum is affected by concentration and temperature, and shifts to lower energies as either is increased. Because of the concentration de- pendence, pulse radiolysis is the preferred method for optical studies of isolated solvated electrons in liquid ammonia. The specific conductance of M-NH3 solutions increases with the metal concentration. In very dilute solutions ((10.2 M), the equivalent conductance decreases as expected for a solution that can form ion pairs. One can infer that the cation and electron are truly separated only at concenr trations below about 10.4 M. Other properties such as sus- ceptibility, activity coefficient and density are nearly the same from one metal to another when compared at the same concentrations. The EPR spectrum of a very dilute solution features a very narrow line with a g—value near the free electron value of 2.0023. As the concentration is increased some interaction between the cation and solvated electron can be observed by the paramagnetic shift in the -6- alkali metal NMR. Although several models have been postulated for the species in dilute M-NR3 solutions, only one is in reasonable agreement with the various properties mentioned above. This model proposes the formation of ionic aggregates which retain the basic characteristics of M+ and e; 1 . Species with stoichiometries such as M, M- o v and M2 could form without changing the optical spectrum or molar volume of the solution. This model would then allow for the spin pairing observed in the static susceptibility and EPR measurements as the concentration is increased. Figure 1 gives three models that have been proposed for species of stoichiometry M-. The first is an example of the ion aggregate model and is EPR active. Perhaps the most interesting properties of metal- ammonia solutions are found in the intermediate concentra- tion range. In this range, a transition from non-metallic to metallic behavior is observed as the metal concentration is increased. The phase diagram for sodium ammonia solu- tions shows a liquid-liquid phase separation at 4 MPM sodium at temperatures below 230 K. The more concentrated, less dense metallic phase floats on top of the dark blue, less concentrated electrolytic phase. As the temperature is increased, these phases become miscible. This phase separation has been compared to the liquid vapor phase separation of a metal near its critical point. A review article by Edwards and Sienko [22] discusses the results of supercritical alkali metal studies and the onset of metal- H ’ \ I I I, \ h; \‘ gum-I ° . "" v ‘ x w" .4/ \z. ,I ‘ mnarnwniziuri 000--~\ H ’---- I H \\ H ' H \,' ‘/ @ H u H . \N, ’ ‘ I \N/ \ é \“ I . I H‘ H /o ‘ ' H ’ /-H\ H-\N'- M’ III-(H .2 ,-—— -N M I ' |H PK I / H I H/ H, H . ‘ ’I \ 7 \ ,’ ‘~ A 69' (in /‘~--" \ I EXPANDED ORBITAL ION PAIR WITH OIELECTRON M". Figure 1. Models for Species of Stoichiometry M-O -3- lization. For a very general but thorough discussion on the metal to non-metal transition, the reader is referred to Thompson [19]. As a final note on metal-ammonia solutions,. the com— pounds isolated from concentrated solutions will be discussed. In general, these compounds retain the golden color of the concentrated solution. Li(NH3)4 is the only truly crystalline alkali metal-ammonia compound that has been isolated. The others are of the general formula, M(NH3)6, and contain Ca, Sr, Ba, Eu or Yb. Structures of the hexaamines show that the metal atom is octahedrally situated among the six nitrogens. Powder neutron diffrac- tion spectra show that not all N-D distances are equal in M(ND3)6 and the pseudo trigonal axis of rotation of the ND3 is 13' from being colinear with the M-N bond. Li(NH has 3’4 at least two crystalline forms with a solid-state phase transition at 82 K. Both structures are body-centered cubic. Magnetic susceptibility measurements also confirm a structural change at 82 K and suggest magnetic ordering at low temperatures. The 82 K phase transition was not present 1n L1(ND3)4. B. Alkali Metals in Amines and Ethers The range of solvents available for nonaqueous metal solutions is limited by low solubilities and by solvent reduction [5-8,19,20]. Only with lithium in methylamine -9- can metal concentrations be obtained that are on the order of those obtained in liquid ammonia. For other solvents, concentrations range from millimolar to undetectable. As an example, the solubility of sodium in ethylenediamine is only 2.4 x 10—3 M at 25'C. Table I gives the solubilities of alkali metals in various solvents [23]. The characteristic blue color of dilute metal-ammonia solutions is also observed in metal-amine solutions with concentrations of 10-5 M or greater. The optical absorp- tion spectra of these solutions feature either one or two maxima. One of these bands may be attributed to the solvated electron, as in liquid ammonia. The second is shifted to higher energy, and is strongly metal dependent but only moderately dependent on the solvent. The species responsible for the second absorption band has been identi- fied as an alkali metal anion. This species is not simply one of those shown in Figure 1 but rather a true anion with no solvent molecules between the nucleus and the valence electrons. The optical spectra of alkali metals in ethylenediamine solutions are given in Figure 2 [24]. Except for that of lithium, these spectra all show the metal dependent peak attributed to M-. In the case of lithium, only a solvated electron peak at 7810 cm"1 is observed. The spectra of K, Rb and Cs show the maximum due to M- and a shoulder due to e- Sodium solutions show solv' only the peak of Na—. -10- TABLE I. Alkali Metal Solubilities in Various Solvents. Metal Solvent Solubility Temperature (M11. ('0) Li ethlenediamine 0.29 25 methylamine xLi = 0.15 -- Na ethylenediamine 2.4 x 10-3 25 methylamine >2 x 10‘4 -50 K ethylenediamine 1.04 x 10-2 25 tetrahydrofuran 5 x 10.6 24 Rb ethylenediamine 1.31 x 10’2 25 Cs ethylenediamine 0.054 25 methylamine >2 x 10‘ -5o -11- .mcwfimwnmcmahnum :H macaw: Hamxa< mo muuommm HMOfiumo .N shaman n.o-. A - .Eu-m ON 0. 0. . 0 _ I q .. and .664 s’._ . T - H.0- 00 0.0 . . N0 0 .0 0.. 0N-. 0. ._ W- E:- x -12- Further evidence of the existence of M— is found in the EPR spectra of solutions which contain only the metal dependent optical band. In contrast to the strongly para- magnetic ammonia solutions, these solutions were dia- magnetic with very weak EPR signals. The filled ns2 orbital of M- would satisfy the conditions necessary to give these spectra. Other evidence is found in the oscillator strength, electrical conductivity and solubility studies. The oscil- lator strength has been measured for Na in ethylenediamine (EDA) and was found to be twice that of e;°1v. The conduc- tance of these solutions decrease with increasing concen- tration as expected for electrolytes. Although this behavior indicates ion pairing as with normal salts, the extent is less than expected. This would occur for ions with large closest approach values. The distance for sodium in methylamine (MA) exceeds 6 A, an indication that Na- is a large ion. Solubility studies show that sodium metal will not dissolve in 1,2-methoxyethane alone. However, the equimolar Na-K alloy does dissolve. This can readily be explained by the formation of K+ and Na-. The most conclusive evidence for M- is found in the alkali metal NMR spectra of these solutions. The NMR spectra feature a peak with a chemical shift that corresponds well to the theoretical value of M-. Further background and the results of alkali metal NMR studies are deferred until Part II of this dissertation. -13- Solvation of alkali metals in amines and ethers may be described by the following equilibria: 2M(8) <____ M + M (1) M_ _-—-> M + e- (2) <---- solv solv ____) + _ solv <-—-- M + esolv (3) where Msolv is the "monomer" and probably represents a contact ion pair between M+ and e;°1v [25]. This species has a characteristic EPR signal with alkali metal hyperfine coupling, and is observed to be present only in very low concentrations. To increase the solubilities of alkali metals, organic complexing agents were used. These ligands, L, introduce a fourth equilibrium M + nL M L (4) which drives equations (1) and (3) to the right. Two main classes of complexants have been used for this purpose, crown ethers and cryptands, originally developed by Pedersen [26] and Lehn and coworkers [27], respectively. Figure 3 features a representative molecule from each class. Most commonly, crown ethers are cyclic ethers with alternating -CH2-CHZ-groups and oxygens. Their abbreviated names reflect the ring size and number of oxygens. For -14.. 2,2,2 C rypto nd (C222) (WW 0 O (0‘ .3 pov’ 2 18-C row n- 6 (18-C-6) Figure 3. Cryptand-2,2,2 and l8-Crown-6. -15- example, 18-crown-6 (18C6) is an eighteen containing 6 oxygen atoms. alkyl or benzo groups attached to the ring. are three dimensional complexants that consist ether linkages between two tertiary nitrogens. abbreviation for cryptands is C-m,n,o where m, numerical representations of the number of in each strand between the two nitrogens. have been used not only to increase alkali metal but also as solvents for alkali metals. Sodium to dissolve in 12-crown-4 to form stable, dark tions. Potassium, rubidium and cesium form solutions in 12-crown-4 [28]. Edwards and studied Na— and Rb_ in these two component metal [29]. ethers Crown The hole The crown ethers may also diameter member ring have The cryptands of ethyl The generic n and o are -CRZCH20-groups Crown ethers solubility was found blue solu- metastable coworkers solutions and cryptands have ion selectivity based of 18- 1argely on crown-6 is 18-crown-6 studies of distortion The cavity and The complexation constant as a function of cavity graphically represented in Figure 4 [30]. larger degree of cation specificity than the larger complexation their cavity sizes. large enough to accommodate either Na+ or K+ but shows a preference for K+. X-ray diffraction the complexes of alkali halides show a ring and loss of symmetry as one goes from K+ to Na+. cations, Rb+ and Cs+, are unable to enter the can give metal:complexant ratios other than 1:1. size is Because of their energies, cryptands show a higher crown ethers. The -16- 2107 .0 0 1006 .0 20 m 1505 .A 1.0'I I I I I I Na" Ag*- K" Rb" Cs+ I 1.00 1'20 1.110 1.00 1.00 Cation Radius (A) ‘Figure 4. Log K versus Cation Radius in Methanol at 25°C. - -17- formation constants (log K) for cryptands range from 1.8 to almost 10 in methanol. The maximum for crown ether in methanol is slightly more than 6 for K+1806. Since the cryptands have 3-dimensiona1 cavities they tend to form only 1:1 complexes. However, in some cases, more than one ion can occupy a single cavity. The optical absorption spectra of metal solutions with crown ethers or cryptands present are temperature depen- dent. The absorption maxima shift to higher energies as the temperature is decreased as for dilute metal-ammonia solutions. The cryptand or crown ether concentration also influences the observed spectra. In solvents where the metal is insoluble, addition of complexant gave M- and/or e optical bands. When the concentration of complexant solv solv band exceeded the amount of dissolved metal, the e increased. The conditions needed for this observation are described by equilibria 1-4. The use of complexing agents in nonaqueous metal solu- tions not only provided accessibility to new solvent systems but also opened up the possibility of the formation of new compounds. When a solution of a pure metal in ammonia or other solvent is evaporated, the electrons recombine with the cation to form metal films or powders. The presence of a complexing agent inhibits this reaction because of the energy required to break the cation-ligand bonds. The synthesis and characterization of compounds of this type are discussed in the next subsection. -13- C. Properties of Alkalides and Electrides A golden precipitate forms from a solution of sodium and cryptand-2,2,2 (0222) in ethylamine as it is cooled below -l5'C [31,32]. The precipitate, which was first observed in 1974, was the first crystalline compound which contained an alkali metal anion and has the formula Na+C222-Na_. The parent compound has been fully charac- terized; even the crystal structure has been determined. The packing of Na+C222 and Na- is represented in Figure 5 [32]. The sodium cation is located in the center of the cryptand, which has its three strands twisted to give an antiprismatic arrangement and to allow a closer fit to the sodium cation. These "expanded" cations are closest-packed in ABCABC repeating layers. The sodium anions are located in the pseudo-octahedral holes .of the lattice. The van der Waals contact distances in the crystal suggest that the radius of Na- is about 2.6 A. Since Na+C222-Na- forms isothermally at O'C from sodium and 0222 in solution, the Gibbs free energy must be negative for the following reaction: + 0222 --—-> Na+C222-Na 2N8 (3) (8) (5) (6) Estimates of the enthalpy, free energy and entropy were made based upon a modified Born-Haber cycle [33]. The cycle is diagrammed in Figure 6 and initial calculations -”7\ I ~19- -20- AH'TOTAI. - LI.) “ Mm * NIsI " MiL‘ N m (AH 80b). (AI-I «ML I (AHNNN AHLATTICI Mm A"'Iomz. - I Aflu. - e (9) NIq) ’ N10) I " “ AHIcmMphm LIo) * M'Ig) 0" M“- Figure 6. Born-Haber Cycle for Formation of M+L-N-. -21- suggested that Na+0222oNa— is unstable to disproportiona- tion to sodium metal and cryptand by 22 kJ mol_1, a measure of the error in these estimates, since the compound is stable. Schindewolf and coworkers developed a sodium elec- trode [34-36] which permitted the cell potential of reaction (5) to be measured directly. It was found that the estimated entropy value was 100 J .01-1 K“1 too low, probably due to errors in the estimation of the solvation or crystal entropy [37]. The observed value of A6298 was -7.1 kJ mol-l. Estimates of thermodynamic stabilities were made for several alkalides which utilize cryptands [23] or crown ethers [38]. Cation radii were estimated by Lehn’s method using the ligand thickness so that the Kapustinskii equa- tion could be used to calculate the lattice energies. Because of errors in the method, these stability estimates serve only as guidelines for synthesis rather than as hard and fast predictions. For example, Li+(1204)2-Cs- is cal- culated to be twice as stable as Cs+(12C4)2-Li-. There- fore, if a lithide is desired, a different complexing agent and cation should be chosen. Subsequent alkalides were prepared by one of three different methods [25,39]. The first is the gradual cooling with or without a solvent composition change to precipitate crystals [39-44]. The stoichiometry of the resultant crystals was not necessarily the same as that of the solution. The second method is slow solvent evapora- -22- tion to yield powders and films with the same stoichiometry as the solution [45]. The third method is direct vapor deposition of the metal and complexant onto a substrate [49,50]. The latter method requires either a solid state reaction or a gas phase reaction near the surface of the substrate. By increasing the complexant to metal ratio, electrides may be made by any of these methods. Films for optical absorption spectra of alkalides and electrides were first made by flash evaporation of solu- tions of metals and complexant in stoichiometric amounts. Later, crystalline compounds were introduced directly into the optical apparatus and redissolved [51-53]. Films were made by flash evaporation as before. There exists the pos- sibility that the films do not contain the same species as the crystals that are dissolved to produce the solution. The maxima of the resultant spectra were compared to the spectra of pure alkali metals in solution. Films from Na+C222-Na- solutions yield spectra featuring an absorption 1 band at 15,000 cm- with a pronounced shoulder at 18,900 cm-1. A small but distinct peak at 24,500 cm.1 was also observed. Decomposition of the films resulted in the loss of all spectral features. Films of potassium, rubidium or cesium with C222 were prepared and are considerably less stable. The spectrum of each of these films consisted of a single asymmetric peak and showed neither a high energy shoulder nor an additional high energy peak. The absorp- tion maxima were at 11,900, 11,600, and 10,500 em‘1 for x‘, -23- Rb-, and Cs-, respectively. These locations correspond well to the M_ bands in ethylenediamine (EDA) at 12,000, 1. Na- in EDA solutions has an absorp- tion band at 15,400 cm-1 but no other band or shoulder. 11,200 and 9,800 cm- The similarity between the homonuclear alkalide films and the solution spectra lead to the utilization of optical spectra as an identification tool for heteronuclear alkalides. Films from solutions of K-Na-1806 in methylamine gave a broad absorption band at 13,300 cm.1 with a shoulder at 9,000 e-‘1 [52]. The main band lies intermediate to the expected locations of Na- and K-. The crystals were found to have up to 18X solvent content. This system was resyn- thesized by the author to produce solvent-free crystals with 1:1:1 stoichiometries [51]. Films made from solutions of the pure crystals had a narrower optical maximum at 14,000 cm-1 with no low energy shoulder. Although this compound was presumed to be a sodide, a clear assignment could not be made. The position of the maximum required a 1000 cm-1 red shift of Na- from the homonuclear 0222 film or an 1100 cm“1 blue shift for K-. The maximum position is even further shifted from the 16,700 cm_1 peak observed for films made from solutions of sodium 18-crown-6 in a 2:1 mole ratio. The compound was presumed to be a sodide because of the selectivity of lB-crown-6 for potassium rather than sodium and the greater stability of Na- compared with K—. -24- Other heteronuclear alkalides also have changes in the position of the absorption maxima from those in solution. Cs+(1806)2-Na- has a sharp peak at 14,600 es'l while Rb+1806-Na— has a maximum at 13,800 cm-l. These compounds were also thought to be sodides. In the case of rubidium and cesium with 18-crown-6, two peaks are observed. One covers a 1000 CI_1 range with its maximum at 11,500 cm—l. This band is asymmetric with the center at 12,000 cm_1. The second peak occurs at 9000 cm-1 and was assigned to trapped electrons. Neither Rb+ nor 0s+ can enter the crown ether cavity completely and the anion assignment cannot be made without ambiguity. Shifts in the absorption maxima may be due to cation-anion interaction. Crown ether complexes can have axial symmetry, which would allow exposure of the cation above and below the ring. Cations complexed with 2 crown molecules should be more shielded and may disturb the anion environment less. That this is not necessarily the case is seen with 12-crown-4 sandwich sodides. Dry films of K+(1204)2-Na_ showed two maxima at 13,100 and 17,000 cs’1 [38]. Films prepared from Na+(1204)2-Na- also had two absorption peaks but these were slightly shifted from the heteronuclear peaks at 13,500 and 17,500 cm,"1 [54]. The origin of the higher energy peak is unknown, but it is reminiscent of the shoulder seen with Na+0222oNa_. The cation-anion interaction should be less for cryp- tated cations than for those salts with the cations -25- complexed by crown ethers. Since in the former case the cation is complexed by a three dimensional complexant, one would expect less interaction of the cation with its surroundings and smaller shifts in the absorption maximum location of the anion for these salts. However, this is not the case as seen in the M+0222-Na_ compounds. The peak location for K+0222-Na— is red-shifted 300 cm"1 from that of Na+0222-Na- while in Rb+0222oNa_ the absorption is shifted to 14,000 em’l. The first optical absorption spectrum of an electride was that of a film made from a solution of potassium and cryptand-222. The maximum at 7400 cm"1 was in the near infrared, as expected for a trapped electron. As the potassium/cryptand mole ratio increased from 1 to 2, a second band corresponding to K_ appeared. At a mole ratio of 2 only the K- peak was observed [55]. Since that time a number of electride films have been prepared. The observed spectra fall into one of two categories -- localized or "metallic" behavior. The localized electrides have absorp- tion maxima which correspond to traps of at least 0.4 to 0.6 eV depth. These absorption bands tend to be broad, indicative either that electron-electron interactions may be important or that the excited state is delocalized. These electrons may also undergo fast exchange with adjacent trapped electrons but not on the optical time scale. The "metallic" electrides give absorption spectra similar to those of concentrated metal-ammonia solutions. -25- The absorbence remains high in the infrared, out to at least 3000 cm-1. This suggests electron delocalization or at least shallow traps. This behavior has been observed from certain K+0222-e— films and films from solutions with stoichiometries of Li20211 and 032 1806. It has not been possible yet to verify the "metallic" behavior of crystal- line electrides by D.C. conductivity measurements because of their instability. However, powders of these stoichio- metries have large microwave power absorptions that would be expected for conducting samples. The golden appearance of Na+0222-Na- suggested that it may be metallic. However, a plot of log resistance versus 1/T had a slope that corresponded to a semiconductor with a band gap of 2.4 to 2.5 eV. Extrapolation to infinite tem- perature gave a specific conductivity greater than 106 ohm"1 cm-l, an indication that Na+0222-Na_ is an intrinsic semiconductor. D.C. conductivity measurements of pressed powders indicate that all alkalides behave as semi- conductors. Reported band gaps range from 0.89 to 2.6 eV. However, most alkalides have much lower specific conduc- 2 to 10 ohm.1 c111-1 tivities of 10— at infinite temperature than expected for intrinsic semiconductors. The observed electrical conductivity is probably due to defect electrons which are trapped at energies much closer to the conduction band rather than from the valence electrons in M_, thereby giving extrinsic semiconductor behavior. Electrides, prepared as powders by solvent evaporation, also behave as -27- semiconductors with band gaps of 0.5 to 0.8 eV. The con— ductivity of electride samples may also arise from "defect electrons" rather than from the bulk of trapped electrons in the sample. The first crystalline electride, Cs+(1806)2-e-, had a specific conductance intercept at infinite temperature of ~102 ohm.1 cm”1 with a band gap of 0.9 i 0.1 eV. Therefore, Cs+(1806)2-e- has characteristics that lie between extrinsic and intrinsic semiconductivity [42,53]. The ns2 electronic configuration of alkalides predicts that they are EPR inactive. Samples of Na+0222-Na_ are indeed diamagnetic and have no appreciable EPR signal. Some samples of K+1806-Na- [51] and Rb+1806-Na- [53] showed comparable strong EPR signals with the presence of fine structure. This structure is due to the hyperfine splitting of the signal by the interaction of trapped elec- trons with the K+ or Rb+ nucleus. The presence of hyper- fine splitting is preparation-dependent and is not evident in all samples. In most cases the EPR spectrum of an alka- lide sample consists of a single, narrow line near the free electron g-value. Sometimes a second broader band is superimposed on the first, indicating the presence of two or more trapping sites. In general, the EPR spectra of alkalides appear to arise from trapped electrons rather than from intrinsic structural features of the crystals. In contrast to alkalides, most electrides have very strong EPR signals. Powders of Li+0211-e— had microwave -28- power absorption similar to those of powdered metals [56]. The 0s+(1806)2-e_ EPR spectra featured a single Dysonian line at g = 2.0023. The line became more symmetrical at lower temperatures. The width of the line was only 0.48 G, which indicated well-defined local trapping of the electrons [53]. Samples of crystalline K+0222-e- and Rb+0222-e- gave such strong EPR signals that only a very few crystals would saturate the detector of the spectro- meter [50]. Magnetic susceptibility measurements of alkalide crystals also show that alkalides are diamagnetic. Some paramagnetism of defect electrons is usually evident at low temperatures. The electrides shows a wide range of magnetic properties. Cs+(1505)2-e- has low temperature deviations from the Curie—Weiss law that allow it to be classified as an antiferromagnetic material [57]. Crystalline K+0222-e- is best described as a Pauli para- magnetic compound. At low temperature there is some adherence to the Curie law but the susceptibility quickly goes to a metal-like temperature-independent value as the temperature is increased. Rb+0222-e- behaves very much like the potassium electride but the Curie-Weiss behavior extends over a larger temperature range. The electrical conductivities of these two salts were too high to measure by the available methods, a further indication that K+0222-e- and Rb+0222-e- are either metallic or near the metal-nonmetal borderline [50]. -29- The object of Part I is to describe the synthesis and characterization of new alkalide and electride salts. Characterization techniques include elemental analysis and optical absorption spectroscopy as identification tools and D.C. conductivity measurements for all compounds. Magnetic susceptibility measurements and EPR spectra were recorded for those compounds thought to be electrides. The results of these studies for 10 new alkali metal—crown ether systems are reported. PART I CHAPTER TNO -- EXPERIMENTAL A. Reagent Preparation 1. .Mbtals - Sodium, potassium and rubidium under argon were purchased from Alpha Ventron Products in 5 gram break- seal ampules. Total purity was 99.95% for each metal. Cesium metal was obtained as a gift from Dow Chemical Corp. and had been previously transferred into ~5 gram sealed glass ampules. Each metal was further distributed to smaller tubing following the procedure outlined by less [52]. Desired quantities of metal were obtained by isolating calculated lengths of metal inside of tubing that had premeasured inner diameters. 2. 15-Crown-5 - (1505 or IUPAC 1,4,7,10,13-pentaoxacyclo- pentadecane) was distilled at 85-90'0 under dynamic vacuum. The apparatus for the distillation is given in Figure 7. The crown ether was pipetted into the bottom of the apparatus without further purification. The apparatus was evacuated to ~2 x 10-5 torr before submerging the lower half into a preheated oil bath at the desired temperature. The cold finger was cooled by flowing dry nitrogen gas -30.. -31_ To Vacuum Manifold / Cold N2 Gas ‘I .\ Trap :0 protect Vacuum Manifold t: Figure 7. cu I p 0" Crow n Vacuum Distillation Apparatus for lS-Crown-S Purification. -32- at 0’0. The temperature of the nitrogen should be warm enough to avoid condensation of any unwanted impurities (such as H20) onto the cold finger. When the crown ether had completely distilled into the bucket, the apparatus was closed, removed from the oil bath and allowed to achieve thermal equilibrium. The entire apparatus was brought into an inert atmosphere box and the crown was transferred with a disposable pipet from the bucket to a storage bottle. 3. Solvents - a. Dimethyl ether (MeZO; Matheson) and tri- methylamine (TMA; Matheson) were first stirred over calcium hydride at -20'C for several hours. The solvent was frozen in liquid nitrogen, pumped to <4 x 10-5 torr, thawed and stored at -78'C overnight. The solvent was then transferred to flasks containing an excess of NaK3 alloy and about 0.5 gram diphenyl ketone, frozen, pumped and stored overnight, before transferring to another NaK3 bottle. After the solution maintained the characteristic blue color of the benzophenone dianion-ketyl mixture for 2 consecutive cycles, the solvent was transferred to heavy walled glass storage bottles or stainless steel storage tanks. The solvent was then subjected to several freeze- pump-thaw cycles until the initial pressure at pumping was less than 1 x 10_4 torr. b. Diethyl ether (ethyl ether, anhydrous, Mallinck- rodt, Inc.) was treated in the same manner as dimethyl ether and trimethylamine. Final storage was over NaK3 alloy without diphenylketone. -33- B. Synthesis Techniques 1. Glassware - All glassware was first rinsed with a hydro- fluoric acid cleanser solution (53 HF (28M), 33% HNO3 (16M) and 628 deionized water, by volume) then immediately rinsed six times with deionized water. The glassware was filled with freshly prepared aqua regia (3 H01:l HN03, by volume) and left overnight. The aqua regia was discarded and glassware was rinsed six times with deionized water followed by six rinses with conductance water (house deionized water which had been further deionized with a Crystalats Deeminizer and distilled through a high reflux ratio column to less than 1 ppm impurity) and dried in a 450°F oven overnight. 2. Improved synthesis technique - The general synthesis scheme for the preparation of alkalides and electrides has been fully described [51]. This section is used to point out the improvements made over this general scheme by the use of a simpler apparatus called an H-cell (Ace Glass) and a new solvent system. The H-cell, shown in Figure 8, was modified from the factory issue to better accommodate the low temperature synthesis of alkalides and electrides. First, the angle between chambers A and g was increased from 90' to about 120' with respect to the horizontal con- nection. Second, the distance between the stopcock and horizontal was increased by 7-8 cm to allow complete Figure 8. Synthesis Apparatus: H Cell. -35- cooling of the frit without cooling the stopcocks. Lastly, a 9 mm OD side arm was added to one of the chambers for metal introduction. The size of the H-cell is its main advantage. H-cells are small enough to allow entry into the inert atmosphere box so that the complexants are never exposed to atmosphere but can be loaded into the synthesis apparatus under the dry helium atmosphere. Syntheses involving lithium are also greatly simplified by direct loading of the lithium in an inert atmosphere box [50]. The 15-crown-5 was loaded through the stopcock into chamber A with a disposable pipet. The amount of crown ether was determined by mass difference of the storage vial. A premeasured amount of metal, sealed in a 3 or 4 mm OD scribed glass tubing was broken and placed in the metal sidearm. The H-cell was closed and sidearm capped with a 3/9” Ultra-torr Union (Cajon) and a closed end glass cap. The H-cell could then be removed from the dry box and evacuated to 2 x 10-5 torr. The Ultra-torr union and glass cap were removed by making a vacuum seal on the sidearm. A metal mirror was made in chamber B by gently heating the remaining sidearm until all metal had distilled into B. A final seal was made at the constriction and the H-cell allowed to thermally equilibrate. The use of dimethyl ether instead of methylamine improved the stability of the alkalide and electride solu— tions. Working temperatures were increased from -40'0 with MA to -10' to 0'0 with MeZO. This allowed for faster dis- -35- solution of the alkali metal and generally shorter synthe- sis times. Chamber A of the H-cell was cooled to <-30'C and ~10 ml dimethylether was distilled onto the now frozen 15-crown-5. The apparatus was closed and warmed to -10'C in a dry ice-isopropanol bath. When the crown was completely thawed and dissolved, the solution was poured onto the metal mirror in A. Me20 was distilled from A to A and then poured back into A several times to insure complete solution of the crown. The metal mirror was dissolved by gentle agitation of the H-cell. Complete dis- solution took 0.5 to 3 hours at -10'C. When no trace of the metal mirror could be seen, the entire solution was poured into chamber A. The total volume of the solution was reduced to a small viscous puddle (~l-3 ml) by distil- ling the excess Mezo to a waste bottle. The cell was cooled to less than -30'C. If there was any precipitation of material, more MeZO was added to yield a homogeneous solution. Diethyl ether (DEE) or trimethylamine (TMA) were added as co-solvents for crystallization. It is important to keep the DME solution temperature well below the DME boiling point of —27'C to avoid reverse distillation. The co-solvent was distilled, a small portion at a time, with the H-cell being shaken between additions to insure proper mixing. Total volume of co-solvent did not exceed 7 ml. The H~ce11 was then packed in dry ice for 3 to 12 hours. -37- The H-cell was then checked for the formation of crystals. If no crystals were present, the solution was further concentrated by removal of the solvent to a waste bottle. The H-cell would again be stored at -78'C for several hours. After the formation of crystals had stopped, the mother liquor was poured into A and then distilled into a waste bottle. The wash solvent for the crystals was the same as the co-solvent. Fresh solvent was distilled onto the crystals, the cell shaken to insure mixing and the solvent decanted to chamber A. The solvent was then distilled from A to A and decanted to A to remove any soluble impurities from the crystals. This distill-decant procedure was repeated several times until the solvent became colorless. Finally the solvent was distilled from A to a waste bottle until the H-cell was solvent free. The crystals were further dried by dynamic pumping to 2 x 10_5 torr. If the crystals had sufficient thermal stability, the H-cell was allowed to warm to ambient temperature for several minutes while pumping. During pumping, the cell was repetitively shaken and tapped to free crystals from adherence to the walls. Crystals were harvested by transferring into either evacuable storage tubes or glass vials. The most stable compounds could be taken into the inert atmosphere box and stored at -40'C in the freezer that is located in the box. Compounds with greater temperature sensitivity were transferred to evacuable storage tubes in a nitrogen glove -38- bag. The storage tubes were evacuated and individual sample fingers were sealed off. These compounds were stored at -80'C. Some syntheses were made in the more complex synthesis " which features an additional apparatus called the "cow sample storage tube chamber. This apparatus has the advan- tage of keeping the crystals under vacuum until characteri- zation studies were performed. The synthesis procedure is identical to the H-cell preparation with the exception that the crystals are moved to the storage chamber prior to washing. After the crystals are washed, the wash solvent is frozen in one of the other chambers. The apparatus is evacuated to (4 x 10.5 torr, as in the case of the H-cell, and the storage chamber is sealed off. A detailed discus- sion of the final stage has been given elsewhere [51]. 0. Analysis The general analysis scheme for determining the stoichiometry of alkalides is given in Figure 9A. The crystals were first allowed to react with water to yield a basic residue and hydrogen gas. The volume of the hydrogen is directly proportional to the reducing power of the crystals. A calculated excess of H01 (at least three times the number of moles of 0H_) was added to the decomposed material. The resultant solution was divided into 3 par- tions; one for flame emission spectroscopy, one for pH -39- A. Mt (I505),- N" + 2 14,0 i / M" “505). 1‘ N+ 4' 20H" * Ham .+ 'icnIsunun i I I . 1 dried flame pH titration Quantitative emission with base 'H NMR / M" 0505);? + I-|20-">M"(|5<35)u+ OH- I V" Ham Figure 9. Analysis Scheme for Alkalides (A) and Electrides (B). -4o_ titration with base and the third for quantitative proton NMR measurements. Figure 98 gives the equation for the reaction of an electride with water. Note that the amount of hydrogen gas evolved is exactly half the amount of elec- tride used. The same analysis scheme was used for elec- trides. This procedure gives three checks for the metal content and one for the crown content of the crystals. 1. HYdrogen evolution - Care was taken that the crystals decomposed only according to the equations given in Figure 9. Sample tubes were scribed with a glass knife and weighed quickly to avoid thermal decomposition. The hydro- gen evolution apparatus is shown in Figure 10. Crystals were loaded under a dry nitrogen atmosphere by breaking the scribed sample tube open and placing both pieces through the 1/2" tubing at the top of the apparatus. The bottom surface was precooled in liquid nitrogen to avoid thermal shock of the crystals. The 1/2" opening was then capped with a sealed glass tube and an Ultra-torr union. The apparatus was evacuated to ~10—5 torr on the vacuum system designed for hydrogen collection. This system consists of a gas buret and a modified Toeppler pump and is described by Van Eck [23]. The entire system had been previously evacuated to less than 2 x 10—5 torr. Conductance water which had been degassed several times was used for the de— composition reaction. The water was allowed to distill through the hydrogen collection system and freeze on the walls above the crystals. The vessel was warmed spot-by- -41.... Ca|on Ultra-Torr Unlon Figure 10. Hydrogen Evolution Apparatus. -42- spot to melt the ice and allow only a small amount of liquid water to contact the crystals at any time. After the reaction began, a dry ice—ispropanol bath was used to cool the vessel in order to slow or halt any further reac- tion. By repeating this thaw-quench procedure, the crystals were slowly decomposed without thermal decomposi- tion to yield a white residue. At this time, additional water was rapidly distilled onto the decomposition product to form a solution and insure complete reaction. The evolved hydrogen was pumped through a double liquid nitro- gen trap into a chamber of known volume by means of a manual Toeppler pump. After several cycles, the pressure of the hydrogen was measured by noting the height of the mercury in a U tube opened on one side to atmosphere and the other side to the known volume of hydrogen. The end point was determined to be the pressure at which five addi- tional pump cycles did not change the mercury level. By noting the temperature of the hydrogen chamber and the atmospheric pressure, the ideal gas law could be employed to calculate the males of hydrogen evolved. The pump cycles also served to evaporate the basic solution to yield a dry residue. 2. pH" titration - The residue from the hydrogen evolution was dissolved in a known amount of H01 and conductance water under a nitrogen atmosphere, to give a solution of pH ~3. The number of moles of hydrogen evolved was used to estimate the number of equivalents of base present. The -43- solution was divided into 3 portions, as described above. The pH titration portion was further divided into 2 or 3 parts. Each part was titrated with an NaOH solution that had been freshly standardized with H01. Phenolphthalein was used to indicate the titration endpoint. To prevent CO2 absorption by the base, the titration buret had a glass sheath allowing dry nitrogen gas to flow over the solution. The results of the different titrations were averaged to give the number of equivalents. 3. Flame emission - A Jarrel-Ash Atomic Absorption/Flame Emission spectrometer was used to determine the alkali metal content of the compound. Only the flame emission feature was used. The flame emission portion of the resi- due solution was further diluted to give solutions of ~35 ppm for each metal. Standard solutions of each metal in the 10-100 ppm range were used to obtain the calibration curve. The reading obtained with conductance water was noted between each standard in order to give a baseline correction. Calibration curves of ppm vs digital output of the amplifier circuit were used to find the concentration of the unknown. 4. Proton MMR‘~ The third portion of the hydrogen evolution residue solution was neutralized by adding the appropriate amount of the NaOH solution. The solution was covered with tissue and allowed to air dry at room temperature. When dry, the resultant residue was further dried in a vacuum desiccator for 30-60 minutes. More rigorous pumping -44- resulted in evaporation of crown and led to errors in the early determinations. A known amount of dry sodium acetate and ~0.5 ml 020 were added to the residue and the solution was transferred to a 5 mm OD NMR tube. The sodium acetate served as an internal standard. A 1:1 ratio of acetate protons to crown protons was approximated by using the amount of hydrogen evolved as an estimate of the number of moles of 15-crown-5. Spectra were taken with a Bruker 250 MHz Fourier Transform NMR spectrometer equipped with an Aspect 2000 computer. Individual peaks were isolated and fitted to either a Lorentzian or Gaussian curve by a line fitting program provided by Bruker [58]. This program gives the amplitude, full width at half height and the standard deviation for the frequency of the fitted curve. A ratio of the area under the curves gave the ratio of acetate to crown protons. D. Optical Spectra Optical transmission spectra were taken of films made from solutions of crystals dissolved in methylamine. Figure 11 shows the apparatus used for optical studies. Crystals were loaded through the stopcock into the pre- cooled optical cell under a dry nitrogen atmosphere. The apparatus was closed and evacuated to 2 x 10-5 torr. Methylamine was distilled onto the crystals to fill the op- tical cell. Films were made by two methods. The first -45- Graded Seal C. H Quartz 4 Optical * Cell Figure 11. Optical Absorption Apparatus. -46.. method consists of pouring the solution from the cell to the side chamber which is held at liquid nitrogen tempera- ture. The solution must be concentrated enough to leave a good film with one pour. The resultant films were fairly uniform to the naked eye. The second method is flash evaporation. The solution was poured into the side chamber leaving 1 or 2 ml in the optical cell. The side chamber was frozen in liquid nitrogen while the apparatus was vigorously shaken to spread the solution over the walls of the cell. These films were not as homogeneous as those made by the first method but less concentrated solutions could be used. Spectra were recorded on a Beckman DK-2 double beam recording spectrophotometer that had been modi- fied to permit temperature control of the sample compart- ment between 0' and -l30'C. Rough temperature control was achieved by adjusting the flow rate of nitrogen gas that had passed through a cooling coil immersed in liquid nitro- gen. A copper-constantan thermocouple was placed in the cell compartment to monitor the temperature. Spectra were recorded in the 2800 to 400 nm range with the reference beam passing though air. The spectra were normalized by subtracting the baseline correction of an empty cell and then rescaling the lowest absorbence to zero and the maxi- mum to 1.0. Temperature dependence of the optical trans~ mission spectra was also studied. -47- E. Pressed Powder Conductivity Powder conductivity measurements were made under pres- sure in an apparatus designed by J.L. Dye and M.R. Yemen [20,59]. Under a dry nitrogen atmosphere, a sample was loaded into a precooled 2 an inner diameter heavy-walled fused silica tube which rested on a stainless steel elec- trode. A second stainless steel electrode was inserted into the top of the tube and was used to compress the sam- ple by means of a spring with a measured force constant. The sample cell was loaded into a cryostat which was cooled by controlled liquid nitrogen boil-off. Ohm’s law obedience was checked by measuring the current as a function of voltage. At a constant voltage, the crystals were slowly cooled to less than -50'0 and warmed again to 0'0. The current was read at. approximately two degree intervals during the temperature cycle. The apparatus was then disassembled to check for decomposition. If the crystals had maintained a dark blue color, the system was reassembled and taken through another temperature cycle. For some samples a check was made for polarization by reversing the current flow. After completion of the second temperature cycle the height of the pressed powder was measured so that the specific resistivity could be estimated. -43- F. Electron Paramagnetic Resonance The samples for EPR spectra were loaded into evacuable 4 mm OD quartz tubes in a dry nitrogen atmosphere. Each tube was evacuated to 4 x 10-5 torr and sealed. Spectra were recorded on an X—band spectrometer (Bruker model 200). Temperature regulation was unavailable except by reduction of the flow rate of cold nitrogen gas. To avoid condensa- tion of water onto the sample tube, the gas flow was main- tained at a high rate. Spectra were taken at -105 to -90‘0. A copper-constantan thermocouple with a digital readout (Doric model DS-350) was placed under the sample tube to observe the temperature. C. Magnetic Susceptibility The magnetic susceptibilities of the electrides were measured by an S.H.E. Corp. computer controlled variable temperature susceptometer equipped with a superconducting quantum interference device (SQUID). Samples were loaded under a dry nitrogen atmosphere into thin wall Delrin buckets with snap-on lids. The buckets had an overall height of 10 mm and 6 mm outer diameter. The buckets were stored in liquid nitrogen until loading into the airlock of the susceptometer. The airlock was evacuated to ~500 millitorr and flushed with helium gas. This evacuate-flush process was repeated 2 or 3 times and the sample bucket was -49- visually checked to be sure no frost remained before the sample was lowered into the SQUID at 5 K. The magnetic moment of each sample was measured at various temperatures between 1.7 and 240 K. The apparent ”susceptibility", x’, was given as the magnetic moment in emu divided by the field strength in gauss. Ten readings were taken at each temperature. The software automatically discarded any data points that had more than a preselected percentage difference between two successive moments. This value was usually chosen to be 10%. After a complete temperature cycle, the sample was ejected and allowed to thermally decompose in the helium atmosphere of the air lock. The apparent susceptibility of the decomposed sample was measured at the same temperatures. Mass determinations of the decomposed sample and empty bucket were made so that the molar susceptibility, x", could be calculated. If the sample mass was known prior to the measurement of the moments, the S.H.E. computer could calculate xM directly. However, the weighing procedure introduces the chance of premature decomposition of the sample. The electronic contribution to the molar suscepti- bility was then calculated by the following equation: I _ I z xsample-tbgcket xdecomposed sample+bucket (6) xM moles of sample PART I CHAPTER THREE -- RESULTS AND DISCUSSION The utility of any new compound is dependent on the ease of preparation and total cost. Electrides and alka- lides are excellent reducing agents with one or two eleC+ trons readily available to donate. Although the environ- ment must be rigidly controlled to avoid decomposition of these salts prior to reaction, the use of alkalides and electrides as electron donors for difficult reduction reac- tions seems feasible. Some of the problems encountered in the economical utilization of alkalides and electrides are the length of the synthesis procedure, decomposition of the solutions during synthesis, the cost of complexants and the thermal stability of the crystals. The major goal of Part I of this dissertation is to describe the synthesis and characterization of new alka- lides and electrides. A new solvent system was implemented for the synthesis, and the synthesis procedure was modified to reduce total time. The use of dimethylether as a dissolution solvent greatly increased the stability of the solutions so that fewer reactions decomposed. A new complexant, 15-crown-5, is relatively inexpensive compared -50- -51.. to most crown ethers and to all cryptands was tried. The success achieved with 15-crown-5 as a complexant is reported in this chapter. In all, 10 crystalline systems were isolated and characterized. A. Synthesis Due to the ease of preparation and thermal stability of Na+0222-Na-, the first system to be attempted with 15- crown-5 was Na+(1505)2-Na_. The sodium cation, with a radius of 1.02 A, is slightly too large to enter the cavity of 15-crown-5 which is estimated to be 1.7 to 2.2 A in diameter. Several syntheses attempts were made with crown ether to sodium mole rations ranging from 0.88 to 1.14 in the hope that we could mimic the l8-crown-6 systems. Cesium forms "sandwich" complexes with 1806 in a 1:2, Cs+ to 1806 mole ratio [53] whereas rubidium forms compounds with metal to crown ether mole ratios ranging from 0.5 to 2.0. Both cations are too large to enter the l8—crown-6 cavity. In the case of sodium with 15-crown-5, no crystal- line precipitates could be isolated from any of the synthesis attempts. One problem was difficulty in dissolving sodium metal in dimethylether (MeZO). Even in the presence of lS-crown-S, dissolution took from several hours to two days. Methylamine (MA) was used either alone or with Mezo to decrease the dissolution time. Although, the solubility of sodium was increased in the presence of -52- MA, the solution stability decreased. Lithium was used to give solution stability in those synthesis in which methyl- amine was a dissolution solvent. The second problem was the precipitation of metal at the time of the addition of the crystallization solvent. Although the presence of lithium reduced the chance of decomposition, its presence compounded the metal precipitation problem. In one synthesis which utilized MA and lithium, long, flat, golden crystals were observed. These crystals appeared to be highly reflective and were reminiscent of the K+1806-Na— crystals which had ~14X solvent content [51]. The sodium 15-crown-5 crystals decomposed before they could be isolated and analyzed. The next system attempted was K+(1505)2-Na_. The potassium cation with r = 1.38 A is too large to fit into the l5-crown-5 cavity even if one uses the largest size estimate. The metal mirror of K-Na dissolved readily in dimethylether 15-crown-5 solution at -10'C to form a dark blue solution. The solution left bronzy-red films on the surface of the reaction vessel. Crystallites began to form immediately when diethylether was added to the concentrated Me20 solution. Precipitation of crystals continued for several hours at 78'C. The crystals are brick red in color and are stable at room temperature for several hours. Five other heteronuclear systems were tried and the attempts resulted in similar success. In addition, preci- pitates formed from potassium or rubidium solutions that -53- had a metal to l5-crown-5 mole ratio of 0.5 or 1.0. Dimethylether was used as the dissolution solvent for all of the syntheses and complete salvation of the metal required 30 minutes to 4 hours. Diethylether (DEE) was initially tried as the co-sovent for crystallization. In these syntheses in which no precipitate formed in the presence of diethylether (DEE), trimethylamine (TMA) was used as the co—solvent. The initial stoichiometry, crystallization conditions and general comments on color and stability of 10 metal 15-crown-5 systems are given in Table II. In general, the precipitates were polycrystal- line or clusters of crystallites and the morphology cannot be described. Several generalizations may be made for these systems. First, the heteronuclear compounds which contain sodium are easier to prepare than those without sodium. Precipitation began with the addition of the co-solvent even at tempera- tures as high as -30°C. The formation of crystals was complete within 30 minutes for the systems with initial stoichiometries of RbNa(1505)2 and CsNa(1505)2. These precipitates were microcrystalline rather than clumps of crystallites. Second, the thermal stability of the com- pounds that contained sodium was greater than that of the other compounds which did not. The precipitate of presumed stoichiometry RbNa(1505)2 was stable for over 24 hours at 25'C. Third, the precipitates from solutions with metal stoichiometry of K1505 and Rb1505 were more stable than _ 4 5 _ .swmoo oa masumhho ho sawusaflaMoaha ham moawsvou unawawpsoon .maasswl oH sow comm an anodes .onsoLA105HA mas: oembt .nason NHtoH .<29 «Amomnvnmmo .mousswl om mom Down as cannon .ousosolvoa ass: comet .nsso: mated .<29 «Amumuvamu .mssoa N ooh comm as cannon .owsosntvos and: vocal .ounvaIIM .mm: «Amomdvnzmo .nasoa an new osmN an vacuum .owsosn mom cecal .ouswvollw .mm: «Amomuvsznm .aaso: N mom comm as ounces .owsoao man: comb: .nsso: m .mu: Anomavom .noasswl om mom cecal as cannon .xosnntosam comet .maso: mum .<29 «Anomavnm .mausswl on now comm an cannon .moslosgm Dewbt .nasos m .oa mom comm us cannon .vos magma comet .ounwvallw .mmn «Amomuvsza .:«8 ca mom assasaonloa loan as magnum .owsoho mas: osmbl ..msa NHIoH .mma Amomuvx .cwa ca sow cecal an cannon .xosdnlosum comet .nsso: mum .uomoo musolfloo adamawvsoo haaoEOMAOMOHm seawuswwaananmso unwawsu . .sovfihuoo~m can mokusaad mlm30h01m~ mo nowawumasam obmusmom can anomawmmoo sowuswmufisuahho .HH mamdh -55- those of the compounds from solutions with a higher 15- crown-5 content. The colors of K(1505)2 and Rb(1505)2 were deep blue-black whereas those of K1505 and Rb1505 were dark bronze. Apparently two distinctive compounds are prepared from solutions with metal to l5-crown-5 mole ratios of 0.5 and 1.0. Fourth, the ease of preparation and relative stability of RbNa(1505)2 indicate that this compound is the best candidate for use as a reducing agent for other reac- tions. B. Analysis Precipitates which form during the synthesis of alka- lides and electrides may or may not have the same stoichio- metry as the solutions from which they form. An elemental analysis scheme has been used which has three checks for metal content and one test for crown ether content. This scheme has been reviewed in Chapter Two of this disserta- tion. Two main problems were encountered with the use of this scheme. First, premature decomposition caused low amounts of hydrogen to be evolved. Decomposition would occur for the more thermally sensitive compounds during the determination of the mass of the scribed sample tube. Great care was taken to prevent decomposition before reac- tion with water but in most cases, the amount of hydrogen evolved was low in comparison to that predicted on the basis of OH—titrations and M flame emission studies. -55- Second, l5-crown-5 is volatile and is removed by rigorous drying of the analysis residue for 1H NMR studies. The crown ether ring can also be destroyed during thermal decomposition and low crown ether content is often encountered. To avoid these problems some samples were analyzed by a shortened scheme. The weighed, scribed sample tube was opened in a nitrogen atmosphere and the crystals were sprinkled into a dilute solution of H01. Portions of the solution were titrated and the rest of the analysis scheme was followed as described in Chapter Two. The large volume of water ensured that the decomposition occurred via the reaction with water rather than by thermal pathways. The results of the analysis of the precipitates are give in Tables IIIA and B. Thermodynamic calculation have been made for compounds of the formula M+(15C5)2-N- and can be used to estimate the relative stabilities of the poten- tial salts [38]. These estimations have been used to pre— dict the species within the compounds from the stoichio- metries. The analytical results have been divided into two categories. The first category is comprised of those results which have a high degree of consistency among the number of moles obtained by the various techniques. The second category contains those compounds whose stoichio- metries can be attributed to a unique compound. Sample sizes were determined by the OH- titration results based on the presumed stoichiometry. The percent deviation from the -57- .N .lsson> amisshv hops: pawns smalsm .-=o mo sowassafla an mosmasoaom sewn sadism n «\s + -so + s-mom-.2 All ohm + -s.s-moe-.z o A so a + -ION + .z + samomav.z All Oman + -z.samomav.z assume mo campuses any so panama 6.51 o.ml m.ct m.¢+ w.m+ m.m+ :22 me NA mIGSOMOImH m.oHI o.H+ m.oHI m.m+ m.m+ mas-s14m m.m+ tn- m.et It- m.e- «.m1 -.-t -.-N1 s.mt It: mahsuxm: .m «.m- It- e.-+ it- m.mt m.-- m.m- m.mm- m.m+ It- m.m+ m.m+ am::OAIou sham clean .2 IINMI mono-owmowoam volsnosm loam sawasw>oa a sea assassa~< «0 usesauns .HHH aaa exams mo N Is. Anomav+nm novsumaasn o>wusaou comm can comm Amomfivnm mad) as owl n a as assuxwa mac~o>ov sovfisosn oomNH N no vsoso N slam. Amomdv+a hso> .xsoa can: coNo~ Anomavomx )sz.-mom-+= asses eeme- «Amos-Vezs (s.N-mom-+s soc-sees Assess :e- eeem- Amos-es us.-mum-s+u sense some . «Amos-Va unassumnn< sawuawhommm lo ddbwusooq zoom asuomow£o«0uw HI vssoaloo .enses as: swan «e soaesam seAsssess< ”sesame we assess: .>H man manna mod N mm.m can x am.m cod x hm.H voH x Nm.m «ca x m¢.m can x mb.m moa x hm.u meg N am.“ nod x hm.N now x mc.m mod N vm.N mod x bc.v ncu x NN.m me“ x mm.N meg x N¢.~ mag x m~.H moa x «H.A voH x mN.N cod x Nm.v mod x mm.~ ooH «IOH v-oH v-cH s-oH nroH area area A-oH ~-°H m-oH ~-c~ ~-o~ m-oH m-cH o-oH o-oH v-o~ «IOH «IoH X NN X XX XX O +I+I+I+I X HOOI-II-I X x HCIN o OOOO +I+LH-H QVDCO XXXX XXXX X C +l +l +l Ncooa c~a3muo firqndm O +I +I +| +l +I +I +I +I 'H s H mmco NNNO Q‘M IO arenas I-ICO mI-IN QQ‘Q'O') NIO -sz.-momsv+sm -o.-eoe-s+se tem.N-moesv+m Ix.~AmomH-+a Isz.mflmom~v+a to.N-momss+m mo.o A mm.o mo.o A ¢A.H mo.c A o-.- mo.e a as.- mo.o A em.~ mo.o A on.“ mo.o w em.o meo.o A so.~ Ne.o A sm.o moc.c a ea.- Ho.o A A-.- moc.o A we.s ~o.e A mo.~ mo.e N am.- mo.c « em.- ao.o A om.~ ma.o H mm.- Nc.c a me.- Ao.o A mo.s we.o A we.“ so «.m vasomloo .aanoIOLQIsOI huw>maosvnoo on Revlon washout ho newsman .> Nflfldh -73- moa x ¢H.N mcu x mc.N mca x mc.~ vo~ x mN.N noH sag xx 1000 I-Isl' I-lco day may vcH voH XXXX N l‘ O b A-cH x m.o a H.H No.o w mm.o «cm x >.o H N.H Ho.c M mm.o N neg x N H N mo.o w em.o (om. Anomav+mo A-o~ x m.o H ¢.v moc.o H mm.o NIoH x N H m moo.° H mm.o N oo~ x H.o w N.H voo.c u m>.o IN. AmomHv+no o-oH x m H m mo.c H om.~ m-oH x w w m No.c a mN.H N ~-o~ x H w m No.9 H bo.H Isz. Amomfiv+mo m-oH x m H m ec.o H m~.~ ~-oH x N H N ec.o w mm.o ~-c~ x >.o H v.m no.9 w om.o N ~-o- x A A s so.o A mm.e use. Anom-+sm ) «8 JWII ago a >0 N .Illfldflddflddll .eumssssoo > manna -79- CD 1'6 .Im.NAmUmHV+Qm Mom B\H msmum> Auswuuso. :4: .ma munmwm .-x.no.x+ b r r h b b D b P — P h b b P n h P P d 4 q d] d d I q d — d d 1 d — d q q d 4 06— OK.— O.—N DAN OfiN OKN OdN U1— (I) -80- gaps. This indicates that the conductivity of these sam- ples is an extrinsic property. Electrons trapped in shallow traps would give rise to the small band gaps and high resistivities. The resistivities at 298 K for these electrides are an order of magnitude lower than those for the alkalides which indicates that there are more carriers in electrides. The band gaps of the electrides are not significantly different from those of the alkalides. The electrides may, however, still be extrinsic semiconductors with the most important carriers trapped in shallower wells compared with the wells for the majority of the trapped electrons. In support of this picture, the optical spec- trum of Rb+(15C5)2-e- for example, also indicated the presence of electrons at many energy levels because of the breadth of the optical absorption peak. There is a general trend for many of the compounds to have an increase in the apparent energy gap as the experi- ment proceeds. This increase of 83 may be due to a "bleaching" of the shallow traps by the current flow or by an annealing or decomposition. The results for Cs+(15C5)2-Na_ indicate that electrons from the shallow traps are removed by the first set of measurements and that the conductivity observed for subsequent temperature runs is due to electrons in deeper traps. There is very little change in the band gap calculated for the second and third data sets. This explanation does not hold for other com- pounds in which the Eg values remain fairly constant, -31- (e.g., K+(15C5)2-e-), or which have erratic values from one temperature run to the next, e.g. Rb+(1565)2-e-. One of the problems encountered in the estimation of the band gap by conductivity measurements is the available temperature range. The conductivity of many materials can be measured over a wide range of temperatures. However, in alkalides and electrides thermal instability of the crystals prevents current measurements above -lO'C. Low current readings limit the low temperature value so that a 100' span is the maximum obtainable range. Slight varia- tions in the energy gap estimation can result in wild deviations for the estimates of resistance at infinite te- mperature. The values of R“ for Cs+(1505)2-Na— differ by a full order of magnitude between two temperature runs even though both have the same calculated energy gap. Data from the KINFIT output show a strong correlation between E and 8 R as expected from Equation (8). Small errors in either Q estimation result in errors for the other. These errors may be plainly observed by the randomness of the R. values. The computer-calculated curves fit the data very well as reflected by the small error in the individual estimates of Eg for each data set. However, the correlation of 88 with R” is too strong and the temperature range too small to use the calculated uncertainty for E8 as the true error of the calculated values. A much more realistic value for the uncertainty of the band gaps should be 10.1 eV or more. -32- In conclusion, all the compounds measured behaved as extrinsic semiconductors with relatively small band gaps. The conductivities for K+(15C5)2-e- and Rb+(1505)2-e— swere higher than those of the alkalide salts which indicates the presence of more carriers in the electrides. The tempera- ture range covered by the conductivity experiments is too small to give accurate estimations of the band gap for such extrinsic semicoductors. It would be be interesting to measure single crystal conductivities for these crystals to see if any ionic conductivity occurs and perhaps obtain more reliable values for the band gaps. E. ‘Electron Paramagnetic Resonance Electron paramagnetic resonance (EPR) spectra give in- formation about the environment of unpaired electrons. Pure alkalides have an ns2 valence electron configuration and would be EPR inactive. Indeed, Na+0222-Na- has no appreciable EPR signal. Paramagnetic impurities, such as F centers or metal atoms, would give EPR signals. Previous studies of alkalide systems yielded EPR spectra of trapped ~electrons that featured hyperfine splitting [51,53]. The splitting of the signal was due to the interaction of the trapped electron with the nucleus of the cation. The amount of interaction and extent of the splitting were pre- paration-dependent for each compound. -33- Preliminary EPR studies were made for K+(l505)2e- and Rb+(15C5)2-e— to determine the line shape and g-value for each compound. Both electrides had very strong.microwave power absorption. A very small amount of sample was used to avoid saturation of the microwave receiver. Even so, power studies showed distortion of the EPR line at power levels of 800 nanowatt or higher. The g-values were calcu- lated from the equation hv = g'p'H (9) where h is Planck’s constant, v is the spectrometer fre- quency, p is the Bohr magneton and H is the field of the resonance in gauss. Rb+(15C5)2-e- and K+(1505)2-e- had very similar spectra and are shown in Figures 19 and 20 respectively. Both compounds have g-values of 2.003 t 0.001 which is very close to the free electron value of 2.0023. The linewidths for the spectra were quite narrow and equaled 0.45 G for n+(15c5)2-e’ and 0.58 a for Rb+(15C5)2-e_ and showed no evidence of hyperfine splitting. The g-value and linewidth indicate that the electron is fairly isolated in its environment and has little interaction with neighboring nuclei that would lead to a g-shift. When the conduction electrons in metals give rise to an EPR signal, the lineshape of the spectrum is highly asym- metric and is referred to as Dysonian. For thick metal plates in which the skin depth, 6, is small compared to the -34- .UOMOHI HM m m.. “momav+nm mo asuuommm mam .mH magmas Tic—Iv. ~85- .oosm- um m.NAmUmHV+s mo asuuomom mam .om museum _...mvu.l+ -86- sample thickness, the A/B ratio exceeds 2.7 [62]. The A/B ratio will have a value between 1 and 2.7 for spherical metal particles in which the skin depth is on the order of the particle size. However, the packing density of pow- dered samples may also affect the A/B value [53]. The line shapes for both K+(15c5)2-e' and ab+(15c5)2-e' were asymmetric with A/B ratios of 2.31 and 1.67 respec- tively. The asymmetry of the line may be due to the intrinsic conductivity of the sample even for nonmetallic samples. For small particles with high conductivity, a Dysonian line shape may be obtained where the skin depth is on the order of the particle size. The A/B ratio usually ranges between 1 and 2.7 for these types of samples. The temperature dependence of the A/B ratio has been used to estimate the apparent band gap of Cs+(15C5)2-e_ [53]. The band gap was estimated to be 0.1 eV from the EPR experiments and 0.9 eV from the pressed powder D.C. con- ductivity experiment. The cause of the difference of E3 between the two methods is not known. Perhaps this difference arises from intergrain resistance or the intrinsic differences in the methods of conduction. The microwave "conductivity" is really a measure of the AC impedance rather than the resistance only and the contribu- tions of both resistance and capacitance components may give an usually small apparent band gap. A number of additional EPR experiments should be done to permit us to understand the cause of the asymmetry of -37- the derivative of the signal. First, temperature studies need to be made to determine if there is a temperature de- pendence of the line shape. Temperature dependence would probably confirm the 0.0. conductivity conclusions that K+(1505)2-e- and Rb+(1505)2-e- are semiconductors rather than metals. Second, the EPR spectra could be obtained for samples which have been diluted with an EPR inactive material. Diluted samples should have Lorentzian line shapes with A/B 1 if the distortion is due to the bulk conductivity of the sample with conduction across the intergrain boundaries. Third, single crystal EPR could be used to study the effects of orientation on the line posi- tion and shape. F. Magnetic Susceptibility Pure paramagnetic materials have susceptibilities that are inversely proportional to the temperature. The propor- tionality constant, C, is called the Curie constant, named after the man who proposed this relationship. For a para- magnetic material with lOOX unpaired spins in the state S = 1/2 and in the absence of orbital angular momentum of the electrons, C = 0.3760. This ”law" was modified by P.R. Weiss to include the mean local field effects generated by ferromagnetic and antiferromagnetic materials above their transition temperature. The resultant relationship is given by -33- - _Q X - T-G (10) and is called the Curie-Weiss law where C is the Curie constant and e is the Weiss constant. The high temperature susceptibility may be used to predict whether a material is ferromagnetic (9)0), paramagnetic (9:0) or antiferromag- netic (e<0). For antiferromagnetic materials, a tempera- ture, T called the Néel temperature, may be approximated N’ by TN = -9. Below the Néel temperature the spins of the material will attempt to align perpendicular to the applied field. If the spins were completely free to move, a tem- perature independent susceptibility would result. However, this is never the case due to crystal field effects or other mechanisms which constrain the motion. In single crystal studies the crystals may be so oriented that the ideal behavior is observed. However, in powdered samples, deviations will be seen from Equation (10) near and below the Néel temperature [63]. Deviations from pure Curie law behavior may also be seen for systems where local interac- tions affect the environment of the electron but in which no distinct transitions occur. Many materials do not have the rigid spin lattice required for them to be classified as ferromagnetic or antiferromagnetic. For these materials, 9 serves as a "catch all" correction parameter. Unfortunately, there is no straightforward way to analyze data that deviate from the simple Curie-Weiss law. -39- The temperature dependence of the susceptibility was measured for Rb+(1505)2-e— and K+(15C5)2-e- up to 225 K. The data were fit, with the use of KINFIT [61], to a modified version of Equation (10) to which a temperature independent correction term, B, was added. Diamagnetic im- purities, such as M—, would give negative values for B and would decrease the observed value of the susceptibility. A correction for paramagnetic impurities, such as precipi— tated metal, can also be made by the use of the B correc- tion term. The plot of l/xM versus T for Rb+(l505)2-e_ is given in Figure 21 for measurements over a temperature range of 1.3 to 224 K. The susceptibility has been corrected to remove contributions from the bucket, crown ether molecules and nonvalence electrons of the rubidium so that only the electronic contribution to the magnetic susceptibility is reported. Two features of the l/xM versus T plot are remarkable. First, at temperatures of 20 K and below, the data deviate from the straight line expected for strict adherence to the Curie-Weiss law. This deviation is in a direction that suggests a decrease in the antiferromagnetic interaction at low temperature. The calculated value of 9, based upon all of the data, was -4.42 x,‘ which indicates that Rb+(15C5)2-e_ exhibits anti— ferromagnetic interactions. The second notable feature of the l/xM versus T plot for Rb+(l5C5)2«e- is the discontinuity in the slope observed between 120 and 140 K. Above 140 K, the plot is -90- SN .Iw.NAmUmHv+nm How wnsumummEmB msmum> EX\H .HN munmwm ONN c_\,.ov..h DON oo— 8— Ov— ON— 8— O. on 8 a? ON ll con LI 8N— -91- linear, but is displaced from the line observed in the 20 to 120 K temperature range. This change was observed as the temperature was decreased as well as increased, which is indicative of a reversible transition. The calculated Curie constant for the entire temperature range was 0.2068. Since the diamagnetic background has been subtracted, we can say Rb+(15C5)2-e- behaves, to first order, as a compound with 55% free spins. The remaining spins may be associated with rubidium nuclei to form Rb-, lie in diamagnetic ground states or be removed by decomposition of the sample. However the deviations suggest that the Curie- Weiss equation is not the correct one to use, so that one should not take the 55X figure too seriously. The correc- tion term, 8, was calculated to be -65 x 10—6 moln1 for this sample. These results might be used to imply the presence of Rb- but, more likely, are a result of an inapproriate equation. Instrument trouble prevented the measurement of the susceptibilty of K+(1505)2-e- below 5 K. As in the case of Rb+(15C5)2-e-, the high temperature data points had good adherence to the Curie-Weiss law but no discontinuity in the slope was observed. The apparent Curie constant is 0.3548 for the electronic susceptibility of K+(15C5)2-e— and corresponds to ~95% "unpaired spins". The experimental points deviate from the calculated line at T g 15 K as seen in the l/xM versus T plot give in Figure 22. The Weiss constant is -19.5 which, if real, corresponds to signifi- -92- .Im.NAmUmHV+M How wwsumummfima mnmuo> EX\H .NN masmfim c_>_ox..w OO— OO 8 O¢ ON 0 O N —b8 3;. -_3 .- O 2 D .- 'IIP uh un- cli- —— uln- an. d- -: III- III- — d u- .- C. ‘- di- -- 4. cu- -- 1 OO. 1094. [Sn 18v 1. com I COO IOOh E -93- cantly higher antiferromagnetic interactions for K+(l505)2-e- than for Rb+(15C5)2-e-. The temperature independent correction term is also much larger and is calculated to be -500 x 10-6 mol~l for K+(15C5)2-e_. The magnitude of this correction term is too large to be attributed to K- or other impurities. Although variations of the B term from one sample to another have been observed in the past, the values of the B term obtained for different samples of the same compound vary by no more than 6 mol-1 a factor of 2 and are on the order of ~40 XIO- [44,50]. The magnitude of the correction term for K+(15C5)2-e- was such that it was first thought to be caused by a change of the zero in the susceptometer between the pristine and oxidized sample measurements. However neither data set, when considered individually, showed evidence of zero drift or discontinuity during the susceptibility measurements. A more plausible cause of the large 8 term is the low temperature susceptibility devia- tions from the Curie-Weiss law. The line fitting program probably overestimates the temperature independent correc- tion terms in order to achieve a good fit with the wrong equation. Careful studies of Rb+(15C5)2-e_ and K+(15C5)2-e- should be made at low temperatures with varying field strengths in order to determine the extent of the anti- ferromagnetic behavior of K+(l5C5)2-e- and Rb+(15C5)2-e-. In addition, the susceptibility of Rb+(15C5)2-e- should be -94- carefully examined in the temperature range of 100 to 170 K in an attempt to verify the existence of the discontinuity. This transition may arise from the ”quenching" procedure which occurs when the sample is loaded into the suscepto- meter at 5 K. Also, more samples of each electride need to be measured to get more reliable estimates of B. Care must be taken that only freshly prepared samples are used for susceptibility measurements due to the apparent decomposi- tion with time during storage at -80'C. G. Data From Other Sources The unusually high thermal stability of the compounds reported in this chapter makes them ideal candidates for a variety of studies. The results of these types of studies, made by investigators other than the author, are included in this section in order to complete the presentation of information gathered for these new compounds. The reader is referred to the publications of the original investiga- tor(s) for full experimental details and data manipulation techniques. 1. giffbrentja] Scanning Cglarigetry —- The thermal stabi- lity of the new compounds that contain sodium is greater than that of the other new alkalides. In general, most sodides are stable up to their melting point (if heated rapidly) and decompose irreversibly at higher temperatures. Two of the new compounds, K+(l5C5)2-Na- and Rb+(1505)2-Na-, -95- were studied by differential scanning calorimetry (DSC) by J.L. Dye at AT&T Bell Laboratories, Murray Hill, New Jersey [43]. The DSC trace for Rb+(l5C5)2-Na- is given in Figure 23. There is a small endothermic transition at 35’C which may be due to the melting of excess crown ether, evapora- tion of remaining traces of solvent or the presence of electride. The compound melts at 75' and then decomposes irreversibly. Na+0222-Na- melts at 73'C and all other alkalides and electrides tested were observed to melt at temperatures below 70'C. Therefore, Rb+(15C5)2-Na- is the most stable alkalide to date. Rapid decomposition begins at ~100°C and is marked by the large exothermic peak at 121'C in the DSC trace. K+(15C5)2-Na- was observed to melt at 45°C with no prior endothermic peak. Decomposition was only 10* at 97'C and reached a maximum rate at 108°C. The thermal stability and ease of synthesis of Rb+(1505)2-Na- and K+(15C5)2-Na_ make these compounds the most feasible for possible utilization as two-electron donors for difficult reduction reactions. 2. gEray Diffraction Studies -- The crystal structure of Rb+(15C5)2-Na- was determined by X-ray diffraction studies of a single crystal. Measurements were made by 0. Fussé and D. Ward with a Nicolet P3F Diffractometer. The rubidium cation is sandwiched between two 15-crown-5 mole- cules as suggested by the stoichiometry. The crown ether rings are staggered such that the Rb+ serves as an inver- sion center. The ten oxygen atoms are more-or-less at -95- .uaz.~ . AmUmH-+nm you mouse own .mm onsmwm cc- 3. ON 3.. wcah<¢wa5mh . - O: 00. OO 00 on 8 - a1]. [a d d — q - u - . - . A d J -. a 10m .. .- -. .. - - . n .- - . ._ .- - no? u a . ' Il‘lIl-u‘inillllul. \\\\ III‘ 1 cm \\ loo ~ N . ~ 18. (Mm) MO‘IJ .LVBH -97- equal distances from the Rb+ and the ether rings lie planar and parallel to one another. The "expanded" cations are stacked in ABAB repeating layers with the crown ether rings tilted a few degrees from the column of stacking. The sodium anions are situated in the pseudo-octrahedral hole with a radius of 2.4 to 2.6 A. The reader is referred to a forthcoming paper for the full crystallographic description of Rb+(15C5)2-Na_. 3. ‘EZAFS and 144;; -- Extended l—ray Absorption Fine Structure (EXAFS) and fi-ray Absorption gear Edge Structure (XANES) give information on the physical and electronic environment of a nucleus, respectively. The K-edge X-ray absorption spectrum of Rb+(1505)2-Na- is given in Figure 24. The spectrum has three features that were used to identify the environment of the rubidium nucleus. First, the location of the X-ray absorption edge can be used to distinguish between two oxidation states of a nucleus. Second, the relative area of the threshold resonances up to 40 eV from the edge may also be used to identify the elec- tronic environment of the nucleus (XANES). Third, the modulations at 40 eV and beyond, after Fourier transforma- tion may be compared to those of compounds with known structures. By the use of EXAFS, structural information may be obtained even when the crystals are too small or of too poor quality for single crystal X-ray diffraction studies. This information is usually limited to neighbors within a 4 A distance from the nucleus under study. -93- PX \J LLLLLLLLIlllLLllllL-LLIIIIIAJ I4750 ISOOO ~I5250 I5500 I5750 IBOOO Photon energy in W Figure 24. K-edge X-ray gbsorption Spectrum of Rh+(15c5)2-Na . - -99- O. Fussé and coworkers [64,65] have studied the rubidium environment of Rh+(15c5)2-e", Rh+(15c5)2-Na“. Rb+(1505)2-Rb+ and be(1505)2. The energy of the K—edge absorption threshold of Rb- is only 2 eV lower than that of complexed Rb+ salts. Therefore, the relative edge posi- tions are not good indicators of the formal oxidation state of rubidium. XANES is used to assign the oxidation states. The area, A, of each absorption threshold resonance ("white line") was normalized to unit edge jump and compared to that of complexed rubidium model salts and known rubidides. The samples were allowed to react with air and the X-ray absorption of the decomposed sample was measured. The results of these studies are given in Table VI below. Table VI. Relative Normalized Areas8 of Rubidium K-Edge Absorption Threshold Resonances. Compound A (pristine)b A (oxidized)? Rb+(l505)2-e_ 36 36 ah+(15c5)2-Na‘ 29 34 Rb+(l505)2-Rb- 18 29 Cs+(15c5)2-Rh’ 1+6 36 K,Rb(1505)2 16 34 8Relative uncertainty is :4 x 10'2 cm. b cm x 100 -100- The area of the white line for the complexed model salts and the oxidized alkalides and electrides ranged from 25 to 36. ah+(15c5)2-e‘ and Rh+(15c5)2-Na' have areas equal to 37 and 29 respectively which indicates that the majority of the rubidium is in the +1 oxidation state in these compounds. This assignment is further confirmed by the small change in the area upon oxidation. 0n the other hand, the white line of Cs+(15C5)2-Rb_ has a very small area of only 1 to 6 which is on the order of that observed for the pure rubidide, Cs+18C62-Rb— (A = 1). Upon oxida- tion the area changes to 36 for Cs+1505)2-Rb- and to 30 for Cs+(18C6)2-Rb- The compound with stoichiometry Rb(1505) was thought to have rubidium in both the +1 and -1 oxidation states. The white line area of 18 lies in between the values observed for pure rubidides and compounds with only complexed rubidium cations. The value of A increased to 29 upon decomposition thereby confirming the assignment of Rb+(15C5)2-Rb- for this compound. It was hoped that the ambiguity over the anion assignment of RbK(15C5)2 could be resolved with rubidium XANES. Again, the area of the white line had an intermediate value of 16 which increased to 34 upon oxidation of the sample, indicative that there is a mixture of oxidation states as in Rb+(1505)2-Rb-. From this result and the stoichiometry, it is evident that there is a mixture of crystals having both potassium and rubidium present as sandwich cations and anions. -101- The crystal structure of Rb+(1505)2-Na— allowed the EXAFS of this compound to serve as the model result. The Fourier transform of the spectrum of Rb+(15C5)2-e- is virtually identical to that of Rb+(1505)2-Na- where the rubidium is sandwiched by two crown ether rings. The EXAFS and stoichiometry' of this compound identify it as a true electride. Likewise, the EXAFS spectra of Rh+(15c5)2-Rh' and the mixture, RbK(15C5)2, are consistent with the presence of both Rb+(l5C5)2 and Rb-. The EXAFS spectrum of Cs+(1505)2-Rb- shows only very weak modulations as observed for krypton gas. Due to the large radius expected for Rb-, the rubidium anion is expected to behave as an isolated ion in the extended X-ray absorption region. Cs+(1806)2-Rb- also has a very weak fine structure such that no structural information may be obtained for the Rb_ environment [64]. At this time EXAFS and XANES have been successfully applied to rubidium alkalides and electrides. Although sodium and potassium EXAFS studies are feasible, the K-edge of these elements lie at energies too low for the available experimental set up. Cesium K-edge X-ray absorption does not yield useful information due to fast core-relaxation effects. Some studies were made for the L-edge of cesium with fluorescence detection. The signal to noise ratio was, however, too low to provide useful information about alkalides and electrides. Although EXAFS and XANES can clearly identify rubidium compounds, another nuclear speci- fic identification tool is needed for the other alkalides and electrides. PART II: Alkali Metal NMR Studies of Alkalides. PART II CHAPTER ONE -- INTRODUCTION The identification of alkalides and electrides has traditionally been based on the stoichiometry and optical absorption spectrum of each compound. However, the species in the films which are used for optical studies, may or may not, be the same as those present in the crystals. Recently, X-ray absorption near edge structure (XANES) studies have been used to identify crystalline alkalides and electrides that contain rubidium [64,65]. Unfor- tunately, due to instrument availability and other experi- mental limitations, this technique cannot be used to probe the environment of the other alkali metals. Nuclear magnetic resonance provides a direct probe of the electronic environment of a nucleus since the resonance frequency of the nucleus is dependent on the shielding by the electrons. This chapter will cover some nuclear properties of the alkali metals, the application of alkali metal NMR to study alkali metal anions in solution, solid state NMR and finally, an example of magic angle spinning (MAS) NMR as an identification tool. -102- -103- A. Nuclear Properties of Alkali Metals Each alkali metal has at least one magnetically active isotope. Table VII gives some nuclear properties of the alkali metals. Most commonly, NMR studies involve the iso- topes 7Li, 23Na, 39K and 133Cs due to their high natural abundances. In the case of rubidium, however, the less abundant isotope is the nucleus of choice. The relative sensitivity is dependent on the cube of the magnetic moment as well as the natural abundance and nuclear spin, so the larger p" for 87Rb determines the larger sensitivity for that isotope than that of 85Rb. The Sternheimer antishielding factor, 1+7“, is a measure of the amplification of the field gradient of the nucleus produced as a result of polarization effects in the electron cloud induced when the atom, as a whole, is exposed to an electric field gradient. Values of 1+Y“ are determined from the fine structure in electron spectra of atoms by the use of quantum chemical calculations similar to those leading to electrical quadrupole moments. The values listed in Table VII refer to the free cations and will change, if the the electronic wavefunction becomes significantly altered by strong ion-ion or ion-solvent interactions. For several of the alkali nuclei the quadrupole moments are large enough to make quadrupole effects dominate the NMR spectra. Linewidths are a reflection of the reaction .bm mososcmom .mlsuo> enema case use saw: was macaw ofiaosusfl onus on» an em: mo gunman :22 caucus one 09 vessmmomfin .Om mososomom A0 ---------..-----------iii}-----------....-----------:1}---..---------.i-------u- ........ hm ~HH ~-c~ x «p.v moo.on vmm.N cofl «\e moons N.m¢ 5H.O NH.O ~¢p.N m.vN «\n amps N.m¢ ~-O~ x mO.H ON.O mvm.~ N.Nh «\m ammo «.mH n-o~ x ¢.m smo.o seem.o mm.m «\s use % m.m~ e-oH x N.w so.o1 mmN.~1 Nfic.o e was w N.NH v-OH x NO.m OO0.0 _ON.O ~.NO «\o goo - H.m ~-OH x ON.O OH.O OHN.N OOH «\m Mme ¢>.o mN.o Neo.ou mmN.m m.Nm «\n “An vb.O n-O~ x m.m v-O~ x m.¢ NNN.O ¢.b H “no Nsx+av .souomH mafi>wuwmsom mason usoaoz . llzm a H .swmm msoaosz wswvaownuwaso .oo>waoaom sodomssvozo .usvloz .oossvssn< mummfiosz omelwoassmum A auguoswnz masssaaz E"l'I'l'I'll||'|l-'ll‘l'llsll'lll'l'1'"-.II'"|--|"""'-"'lll-'||l'"l""I'l'|"|l"|lll"l:' :l|"lll|'I'Il'-""llIl.I-I"Illl|sllIll'llllllll'llIII-"I'IIII"IIIII|I‘IIIl|l|'|'|||"||l"l"l'l'é .muflaol wuflxud «0 mvwahomchm huvuosz .HHD MAO Ilhv-I A 1100 > not: H llD’H O IIO I m II 0 v H Al N o: u u n H MD +3 “'66: IIO-HI u a.: II can IIH "5H IIUU A Made) u Ila-H v u 0 El u a): II .A ll 0 n n u u IUD H "'64: H 0-H Il+hfl I H flea IIH u s-« n U o "HO A n o-u 1» u o I ~/ II a): II .A ll 0 Nucleus 77 29 73 HMPA -61.7 -63.l 1 -60.5 23Na 73 1204 MA EA THF DEA —61.8 -61.9 -62.1 -62.8 -62.9 -109- 78 not observed -103.4 .5 -101.1 39K 73 29 73 EA 12C4 THF ~185.4 -191 -197.2 -213.6 .1 -211.6 87Rb 73 THE -292 -346.4 6 -344.3 13308 at infinite dilution. (aq) elative to M R )Reference 76. )Reference 74. C "a A -110- Several experimental techniques have been implemented to reduce the anisotropic broadening in solids. One of these, which is called magic angle spinning (MAS), will be discussed in this section. The types of anisotropic nuclear interactions that are of interest are magnetic dipolar, chemical shift anisotropy and electronic quadru- polar interactions. Numerous review articles have been written on the NMR of solids and the magic angle spinning technique [84-88]. The truncated dipolar interaction Hamiltonian, for all nuclear pairs i,j in the solid is given by: 3 2 - : 1 . _ 7( Z /2Yirjfi ri.(I. I. 31 2 D iosn mouso>o zoom .sosss-sc reassess so usoosoo .se nose asses Hoosaoso .uononusosmm :« sm>mu hasflmasoosa Ass-cos A-vmm-u seasons asvmm-u assume on Ao-voms As-sm-u Aeneomm Adena-u Aosvoom Asymm-n Ao-vomm A-vem-u AmvN.ONl o Am.vm.ONl Aosvoou An.em-+ mm|4s\ss4 oflddldfll nuluns: Hm.mm-uuuu nonvomm Aavvmal Aomvomv ANvmmal AOMVOO¢ AHvamHl AOMVOOM A~v~m~1 AONVOO¢~ Am.vmb~+ AOHVOOH Am.vbN~+ a: .~\-M4. Alum .w IIIII N22 m¢¢.mmlllll Aomcomm AN-em-l Aomvoom AN-mm-n Aomvomm AH-mm-u Aomvmmm A-cmm-u Ace-cum Am.-me-+ ASH-ems Am.-mms+ mm|4s\ssd mumldws s ...... as: om.mmuuuul mom~nm.m Oomunm .sm.n-eomH-.no lam-NAmUmHVOQD -se.sflmomH-.sm a-momH-ne.u Zomba Ham Home p:=omfi¢o .nosososeous mouse as possess eszam0.5 MHz for K- as discussed earlier in this chapter. If the external field gradient is assumed to be the same for Rb- as for Na_, then correction for the quadrupole moment and Sternheimer antishielding factor (of the cations) yields an estimate of 52.2 MHz for Rb-. The absence of a signal for complexed Rb+ is not sur- prising in view of the large quadrupolar broadening. Based upon the quadrupole~broadened MAS-NMR signal of Na+ in Na+(1806)-SCN_, the MAS linewidth of Rb+(1505)2 is esti- mated to be ~60,000 Hz which is too broad to detect with the available instrumentation. Another method to estimate the linewidth for Rb+Cn is to examine the 87Rb NMR results ~148- from solutions. The linewidth of 150 Hz for Rb” in solu- tion, together with a nuclear quadrupole coupling constant of 1.2 MHz estimated from the solid state spectra, yield an approximate correlation time of ~1010 s for this species. This estimation is extremely rough since the field gradient in solution may differ appreciably from that in the solid. The complexed cation, being larger, probably has a longer correlation time than the anion. However, if the same correlation time is used for the cation as for Rb_, the linewidth yields a quadrupole coupling constant of ~7 MHz for Rb+(15C5)2. This, in turn, would give a MAS linewidth of ~24,000 Hz, which is also too broad to measure. The static and MAS spectra of Rb—, with and without proton decoupling, were measured at 65.4 MHz. The results for four compounds are given in Table XII. The quadrupolar contribution is expected to increase by a factor of 2.7 to 3.6, depending upon the asymmetry parameter, n, from MAS to static spectra [96]. The static coupled and MAS coupled spectrum of Rb+(1505)2-Rb- is given in Figure 32. Since the linewidths observed by MAS-NMR were essentially unchanged upon proton decoupling, the dipolar contributions from coupling to protons must have been removed by spin~ ning. The increase in linewidth from spinning to static- decoupled spectra falls within the range expected from the quadrupolar contribution. This indicates that the contri~ bution from chemical shift anisotropy is small. Results of 37E) I"! with and without Proton Decoupling at 65.4 “12. TABLE XII. 1H Decoupled 1H Coupled 1H Decoupled 6, ppm from Rb’ infinitea lH Coupled MAS Static Static Dilpion und Co ~149- AA CO NM vv O 108 00¢ AA GO 0300 vv OD bCD 00¢ AA OO COM VV 88 OQ‘ F4 F1 AA 00 MM vv GO ON Q‘m -, Fl AA P! F! VV p4 Pi 0'30) 7"? Rb‘(15C5)2-Rb' KRb(15C5)2 490(30) 490(30) 1160(30) 1830(30) ~189(2) Cs’(l505)2-Rb’ 2210(30) 650(30) 680(30) 2450(30) -194(l) Inb- Cs’(1806) aUncertainty given in parentheses. ~150- j J l l ~IOO -200 -300 Chemical Shift from Rb+ (00). ppm 0)- Figure 32. 87Rb NMR Spectra of Polycrystalline Rb+(15C5) °Rb' at 65.4 MHz. Top, static coupled spectrum; bottom, MAS coupled spectrum. ~151- Reduction of the static linewidth of Rb- upon proton decoupling ranged from ~250 Hz for Cs+(18C6)2-Rb- to ~650 Hz for Cs+(15C5)2-Rb-. Calculation of the expected dipolar contribution by the method of Van Vleck [105] would require knowledge of the crystal structures, which are not avail~ able for these compounds. However, an estimate can be made by increasing the known Na- to H distances for Cs+(1806)2-Na- [57] to accommodate the larger size of Rb-. These calculations yielded a dipolar contribution of ~900 Hz which is much larger than that observed for 0s+(1806)2-Rb‘. This suggests, as with the sodides [57], rapid motion of the CH2 protons of the 18~crown~6 molecule at the temperatures at which the NMR measurements were made. The use of 87Rh MAS-NMR provides an excellent confirma- tion of the presence of Rb- in crystalline alkalides. How- ever, the absence of a signal does not necessarily smean that Rbfi is absent. The signal of complexed Rb+ in the solid is too broad to detect but both Rb+(15C5)2 and Rb“ were detected in MeZO solutions. The quadrupole coupling constant of Rb— in crystals is 1.2 t 0.2 MHz, while that of complexed Rb+ is probably at least an order of magnitude larger. Comparison of proton-coupled and proton-decoupled static linewidths indicated the presence of ~CH2 motion in the complexant, even in the solid state. Extrapolation of the chemical shift to infinite frequency gives an isotropic chemical shift of ~186 t 2 ppm for Rb- in crystalline ~152— rubidides, which falls within the range of Rb- chemical shifts in solution. D. 1330s NMR Spectra Extensive 133Cs MAS-NMR studies have been done by Ellaboudy and coworkers [42,53,99] and Dawes [57] as described in Part II-Chapter One of this disssertation. In the present work, 13303 MAS-NMR was used to identify the compounds whose synthesis and characterization were described in Part I. The unusually small electric quadru~ pole moment of ~3 x 10.27 cm2 and 100 percent abundance of 133Cs provide ideal conditions for NMR studies. The chemi- cal shifts of Cs-, Cs+(1806)2-e_ and Cs+(1806)2-X-, where X- is a diamagnetic anion, are separated by several hundred ppm [53]. It was hoped that the peaks of the corresponding l5~crown~5 complexes would be similarly separated from Cs— and each other so that exact identification could be made. The 13303 MAS-NMR spectrum of Cst(15C5)2.5 represented in Figure 33 clearly shows the separation of the various cesium signals. The peaks at +505, +29 and ~263 ppm from Cs+(aq) at infinite dilution are assigned to Cs+(l505)z-e-, Cs+(1505)2-X- and Cs-, respectively. This sample had a 87Rb MAS-NMR peak at ~194 ppm which indicates that Rb_ was also present. Obviously, this sample consisted of a mix- ture of Cs+(1505)2-e-, 0s+(1505)2-Rb' and probably Rb+(15C5)2-Rb_ with some Cs- contamination. Another sam- ~153- MMH . am NAmUmHV m0 m0 mamfimm masmEH as NO Esuuommm mEZImdz mo .mm musmflm . + Ema .33 too 30¢... Pain 440.8qu 2.- 2...- J] 4 — OONI O OON+ 00¢... 80... 80+ - . - . a - - _ - s - - - ~154- ple, without excess 15~crown~5, gave a cleaner spectrum. 133Cs The MAS-NMR spectrum of Cs+(l5C5)2-Rb_ is shown in Figure 34. The major peak at ~29 ppm corresponds to Cs+(1505)2-X-. A small but discernible peak at ~265 ppm indicates a small amount of Cs- in the sample. 133 The results of the Cs MAS-NMR studies of the 15- crown~5 alkalides that contain cesium are given in Table XIII. Each compound has a chemical shift in the range of 24-29 ppm which corresponds to diamagnetic Cs+(15C5)2 salts. Often a small peak at ~260 ppm would be observed in the spectrum of Cs+(15C5)2-K- or Cs+(1505)2-Rb- and indi- cated a small amount of Cs- in these samples. The narrow- ness of the Cs+(1505)2 peak (l30-270 Hz) compared to that of Cs- in Cs+(18C6)2-Cs— (350 Hz) is unexpected for a cesium cation in an environment with axial symmetry. Tem- perature studies by S. Dawes show an increase of the line~ width for Cs+(1505)2 with a decrease in temperature [57]. The unusual narrowness of the complexed cation peak may be due to molecular motion at higher temperatures as in the case of Rb- previously observed. The frequency dependence of 13305 MAS-NMR spectra studied by Ellaboudy indicates that the quadrupole interac~ tions are minimal and that the chemical shift anisotropy and magnetic dipole interactions are the main source of line broadening in Cs+(1806)2 compounds [53]. In contrast to the 87Rb MAS-NMR studies, the chemical shifts were inde— pendent of the Larmor frequency. The quadrupole coupling ~155- s A —‘A__ v—w v—v 'v A~ A v " —— l L l l L l l l J l l l l L 1 +000 +600 +400 +200 0 -200 ~400 ~600 _ CHEMICAL SHIFT FROM Gothic), ppm . 133 + - Figure 34. Cs MAS-NMR Spectrum of Cs (15C5)2-Rb at 23.62 MHZ. ~156- TABLE XIII. Results of 13303 MAS NMR Studies. 6, ppm from Cs‘ at Infinite Compound Dilption AM4/2L_E§ CsI +284 ~~~ Cs’(15C5)2-Na‘ +24 130 Cs*(1505)2-K‘ +24 270 Cs’(15C5)2-Rb‘ +29 250 constants for Cs+(18C6)2 and Cs+(1505)2 were calculated from the location of the satellite peaks with respect to the central transition for each spectrum. 133Cs MAS-NMR is very useful for the In conclusion, identification of alkalides and electrides since the com- plexed cation as well as the anion are easily observed. The chemical shift of Cs+(1505)2 is independent of the anion as long as the anion is diamagnetic. Impurities such as trapped electrons cause large paramagnetic shifts which are indicative of cation-anion interactions. Small amounts of Ca. could be detected in some samples. The chemical shift of Cs_ was ~~260 ppm in contrast to the calculated value of ~346 ppm for Cs-(g) [76]. The large paramagnetic shift of Cs- from the gaseous anion indicates strong inter~ action of Cs- with the environment and the less effective 1330s nucleus by the ns2 electrons than that observed for 23Na and 39K. shielding of the SUMMARY AND SUGGESTIONS FOR FUTURE WORK Ten new crystalline systems were isolated from solu- tions of alkali metals and l5~crown~5. Elemental analysis and optical absorption spectroscopy identified two of these compounds as the electrides, K+(15C5)2-e_ and Rb+(15C5)2-e_. The remaining compounds are alkalide salts. In all cases the alkali metal cation is thought to be "sandwiched" between two 15~crown~5 molecules. This con- figuration has been confirmed for Rb+ in Rb+(1565)2-Na- by single crystal X-ray diffraction studies. The use of opti- cal absorption spectroscopy as an identification tool was found to yield ambiguity in the case of systems with the stoichiometry KRb(l5C5)2 and Cst(1505)2. The assignments of these salts as a mixed alkalide and Cs+(15C5)2-Rb_, respectively were based on rubidium XANES data. In general, only the sodide salts could be identified by opti- cal absorption spectroscopy with a fair degree of cer- tainty; the other alkalides had very broad absorption peaks and it was difficult to make exact assignments of the anions. Pressed powder conductivity measurements indicate that all of the new compounds behave as semiconductors with relatively small apparent band gaps. The calculated resis— ~157- ~158- tivities are much larger than those expected for intrinsic semiconductors with the observed band gaps. The conducti~ vity arises from impurities in the crystal. The electride salts, K+(1505)2-e‘ and ab+(1505)2-e’, both had higher con- ductivities than the alkalides salts, indicative of the presence of more carriers in the electrides. Preliminary EPR and magnetic susceptibility studies were used to probe the environment of the electron in H+(15C5)2-e_ and Rb+(l5C5)2-e—. The EPR spectra of each electride had very narrow, single lines at the free elec~ tron g-value. The narrowness of the line and the absence of a g—shift shift indicate that the electrons have very few interactions with neighboring atoms and undergo rapid exchange. The lines were asymmetric with A/B ratios less than 2.7. This lineshape is consistent with those observed for small particles with high conductivity where the skin depth is on the order of the particle size. Spectra were recorded at only one temperature. The molar susceptibility of each electride deviated substantially from Curie-Weiss behavior at low tempera~ tures. The apparent Curie constants at high temperature correspond to 94 percent unpaired spins for K+(15C5)2-e- and 55 .percent unpaired spins for Rb+(15C5)2-e-. Each electride had a temperature independent contribution to the observed susceptibility. In the case of Rb+(l5C5)2-e-, a small amount of Rb_ may be present. The temperature inde- pendent contribution for K+(1505)2-e— is too large to be ~159- attributed to K A probable cause of the magnitude of this term is the use of an inappropriate equation to describe the magnetic behavior. In general, the new alkalides and electrides are rela- tively stable. Most of the new compounds may be handled at room temperatures for time periods up to 10 minutes. The stability makes characterization studies relatively easy to complete. Differential scanning calorimetry studies deter- mined the melting points of Rb+(1505)2-Na_ and K+(15C5)2-Na- to be 75 and 45’0, respectively. The higher melting point of Rb+(15C5)2-Na- makes it the most stable alkalide characterized to date. The ease of preparation and stability indicates that Rb+(1505)2-Na- is the most likely candidate to be used as a two electron reductant. For the alkalides, the stablity decreases with an increase in anion size. The electrides are sufficiently stable to resist decomposition at room temperature for short periods of time. However, decomposition of the crystals occurs at a very slow rate when stored at ~80'C. Identification of alkalides and electrides without ambiguity has been a major problem. Alkali-metal magic angle spinning NMR is a technique that can be applied to each alkali metal for the identification of alkalides and electrides. The MAS-NMR spectra have been recorded for several alkalides. The generalizations, which can be made from these spectra, are as follows: ~160- ~23Na NMR can be used to detect the presence of both the complexed sodium cation and the sodium anions in the solid. The chemical shift range observed for Na— is ~56 to ~63 ppm, which indi~ cates very effective shielding of the 2p orbitals by the filled 3s orbital. ~Potassium 39 was used for the NMR studies. The first NMR observation of K—, both in solution and in the solid state was made. The chemical shifts observed were ~99.3 ppm for K_ in dimethylether and ~105 to ~115 ppm for K— in crystalline potas~ sides. These shifts are in good agreement with the calculated value of K22) at ~107 ppm and are reflective of highly efficient shielding of inner orbitals by the two 4s electrons. No signal was observed for the complexed cation, presumably due to the large quadrupolar coupling constant. ~87Rb MAS-NMR can detect Rb- in crystalline alka- lides but the absence of a signal does not neces~ sarily mean that Rb- is absent. The signal of complexed Rb+ in the solid was too broad to detect even though both Rb+(15C5)2 and Rb- were detected in MeZO solutions. The chemical shift of Rb- in the solid state was dependent on frequency and extrapolation to infinite frequency yielded an isotropic chemical shift of ~186 r 2 ppm. The ~161- calculated chemical shift value of Rb(g) is ~214 ppm. Therefore, Rb- is slightly more de~ shielded than Na” or K” 133Cs MAS-NMR for ~The chemical shifts observed by Cs+(1505)2-e-, Cs+(15C5)2-X_ and Cs- are suffi- ciently separated so that these species can be readily identified. Therefore, 133Cs MAS-NMR can be used to check the purity of a given alkalide system. Three alkalides were identified by 133Cs NMR. The chemical shift observed for Cs- was ~260 ppm compared to the calculated value of ~348 ppm for CsEg). This value indicates even further deshielding of Cs- relative to Rb-. ~The deviation of the chemical shift of “(s) from the calculated value for MEg), calculated increases with the atomic number for the alkalide anions. This could be caused by less effective shielding of p and/or d electrons by the valence s electrons 95 by an admixture of p and/or d charac- ter with the ground-state 3 wave functions. Finally, seven of the new crystalline alkalides, which were described in Part I, have been identified without ambiguity by alkali metal NMR. They are K+(15C5)2-Na-, x+(1505)2-x', nb+(15°5)2'"a', Rb+(15C5)2-Rb_, 0s+(1505)2-Na', 0s+(15c5)2-x’ and Cs+(1505)2-Rb_. The eighth system, which had the stoichiometry KRb(l5C5)2, was ~162- shown to be a mixture with K‘, nb‘, nb+(1505)2 and K+(1505)2 present. The work presented in this dissertation can serve as a springboard for several other experiments., First, the electrides, K+(l5C5)2-e- and Rb+(1505)2-e—, appear to have fascinating magnetic properties. Careful, low field, low temperature susceptibility measurements are needed to determine the exact types of localized magnetic interac- tions. The EPR studies of the electrides should also be continued to include the temperature and dilution depen- dencies of the line shape. Second, a need for crystal structures is present and these compounds, for the most part, have sufficient stabi- lity so that X-ray diffraction studies are feasible. These studies require high quality single crystals and therefore, the crystallization techniques need to be improved. Besides X-ray diffraction studies, high quality single crystals could be used in investigations of the orientation dependence of the EPR and MAS-NMR spectra and AC and DC conductivity experiments. Third, the presence of a Ca. signal in the 13303 MAS- NMR spectra indicates that the final alkalide in the series, Cs+(15C5)2-Cs-, may be synthesized. Fourth, the field dependence of the K— signal should be checked by 39K MAS NMR at higher field strengths. The availability of appropriate instruments is limited by the low Larmor frequency of the 39K nucleus and these studies ~163- would have to be done at other NMR facilities at this time. Finally, the synthesis of new alkalides and electrides should be pursued with other complexants such as 12~crown~4, 21-crown-7 or "lariat" ethers. The most easily available may be those compounds that contain 12~crown~4 due to its relatively low cost and commercial availability. New electrides are especially desirable for the study of magnetic properties and the understanding of the nature of the trapped electron. 10. 11. LIST OF REFERENCES P.P. Edwards, Adv. Inorg. Chem. Radiocbem. 25, 135 (1982). W. Weyl, Annalen der Pbysik und Chemie £92, 601 (1864). G. Lepoutre and M.J. Seinko, Eds., "Metal-Ammonia Solutions, Colloque Weyl I", W.A. Benjamin, New York, 1964. J.J. Lagowski and M.J. Seinko, Eds., "Metal-Ammonia Solutions, Colloque Weyl II", IUPAC Buttersworth, London, 1970. J. Jortner and N.R. Kestner, Eds., "Electrons in Fluids, Colloque Weyl III", Springer-Verlag, Berlin, 1973. J.L. Dye, Conference Organizer, "Colloque Weyl IV, Electrons in Fluids ~~ The Nature of Metal-Ammonia Solutions", J. Phys. Chem. 79(26), (1975). B. Webster, Conference Organizer, "Colloque Weyl V, The Fifth International Conference on Excess Electrons and Metal-Ammonia Solutions," J. Phys. Chem. gg (10), (1980). J. C. Thompson, Conference Organizer, "Colloque Weyl VI, Le Dernier Colloque Weyl, The Sixth International Conference on Excess Electrons and Metal-Ammonia Solutions," J. Phys. Chem. 88 (17), (1984). C. A Kraus, J. Chen. Educ. §0, 83 (1953). W. L. Jolly, Frog. Inorg. Chem. A, 235 (1959). M. C. R. Symons, 0. Rev. 13, 99 (1959). ~164- 12. 13. 14. 15. 16. 17. 18. 19. 20. 21. 22. 23. 24. 25. 26. 27. 28. 29. 30. ~165- , 165 (1968); Angew U. Schindewolf, Angew Chem. 8 l 68). 0 Chem. Int. Ehg. Ed. Z, 190 ( 9 T. P. Das, Adv. Chem. Phys. A, 303 (1962). J. L. Dye, Sci. Amer. 8A8 (77), (Feb. 1967). M. H. Cohen and J. C. Thompson, Adv. Phys. AZ, 847 (1968). R. Catterall and N.F. Mott, Adv. Phys. A8, 665 (1969). L. Kevan and B. Webster, Eds., "Electron-Solvent and Anion-Solvent Interactions", New York, 1976. M.C. R. Symons, Chem. Soc. Rev. 8, 337 (1976). J. C. Thompson, "Electrons in Liquid Ammonia”, Oxford University Press, Oxford, 1976. J. L. Dye, Prog. Inorg. Chem. 88, 237 (1984). J. T. Lin, G. P. Hagen and J. E. Ellis, J. Am. Chem. Soc. A88, 2296 (1983). P. P. Edwards and M. J. Sienko, Acc. Chem. Res. A8, 87 (1982). B. Van Eck, Ph.D. Dissertation, Michigan State University, 1983. R. R. Dewald and J. L. Dye, J. Phy. Chem. 88, 128 (1964). J. L. Dye, Angew Chem. Int. Ed., Eng. A8, 587 (1979). C. J. Pedersen, J. Am. Chem. Soc. 88, 7017 (1967); 88, 386 (1970). B. Dietrick, J.~M. Lehn and P. J. Sauvage, Tetrahedron Lett. 83, 2885 (1969). R. R. Dewald, S. R. Jones and B. S. Schwartz, J. Chem. Soc., Chem. Commun. A888, 272 (1980). D. M. Holten, P. P. Edwards, D. C. Johnson, C. J. Page, W. McFarlane and B. Wood, J. Chem. Soc., Chem. Commun. A883, 740 (1984). J. D. Lamb, R. M. Izatt, C. S. Swain and J. J. Christensen, J. Am. Chem. Soc. A88, 475 (1980). 31. 32. 33. 34. 35. 36. 37. 38. 39. 40. 41. 42. 43. 44. 45. 46. 47. 48. ~166- J. L. Dye, J. M. Ceraso, M. T. Lok, B. L. Barnett and F. J. Tehan, J1 Am. Chem. Soc. 88, 608 (1974). F. J. Tehan, B. L. Barnett and J. L. Dye, J. Am. Chem. Soc. 88, 7203 (1974). J. L. Dye, J. Chem. Ed. gag, 332 (1977). U. Schindewolf and M. Werner, J. Phys. Chem. 88, 1123 (1980). M. Werner and U. Schindewolf, Ber. Bunsenges. Phys. Chem. 88, 547 (1980). W. Cross and U. Schindewolf, Ber. Bunsenges. Phys. Chem. 88, 112 (1981). U. Schindewolf, L. D. Le and J. L. Dye, J. Phys. Chem. 88, 2284 (1982). . N. Tientega, M.S. Dissertation, Michigan State F ‘University, 1984. J. L. Dye, C. W. Andrews and S. E. Mathews, J. Phys. Chem. Z8, 3065 (1975). B. Van Eck, L. D. Le, D. Issa, and J. L. Dye, Inorg. Chem. 8A, 1966 (1982). D. Issa and J. L. Dye, J. Am. Chem. Soc. A83, 3781 (1982). A. Ellaboudy, J. L. Dye and P. B. Smith, J) Am. Chem. Soc. A88, 6490 (1983). J. L. Dye, J. Phys. Chem. 88, 3842 (1984). D. Issa, A. Ellaboudy, R. Janakiraman and J. L. Dye, J. Phys. Chem. 88, 3847 (1984). J. L. Dye, M. R. Yemen, M. G. DaGue and J.~M. Lehn, J. Chem. Phys. 88, 1665 (1978). M. G. DaGue, J. S. Landers, H. L. Lewis and J. L. Dye, Chem. Phys. Lett. 88, 169 (1979). J. L. Dye, M. G. DaGue, M. R. Yemen, J. S. Landers, and H. L. Lewis, J. Phys. Chem. 88, 1096 (1980). J. S. Landers, J. L. Dye, A. Stacy and M. J. Sienko, J. Phys. Chem. 88, 1096 (1981). 49. 50. 51. 52. 53. 54. 55. 56. 57. 58. 59. 60. 61. 62. 63. 64. 65. 66. ~167- L. D. Le, D. Issa, B. Van Eck and J. L. Dye, J. Phys. Chem. 88, 7 (1982). M. K. Faber, Ph.D. Dissertation, Michigan State University, 1985. M. L. Tinkham, M.S. Dissertation, Michigan State University, 1982. D. Issa, Ph.D. Dissertation, Michigan State University, 1982. A. S. Ellaboudy, Ph.D. Dissertation, Michigan State University, 1984. F.N. Tientega, unpublished results. M. G. DaGue, Ph.D. Dissertation, Michigan State University, 1979. J. S. Landers, Ph.D. Dissertation, Michigan State University, 1981. S. B. Dawes, unpublished results. DISNMR, Disk Interactive Spectroscopy on the Aspect 2000 minicomputer, version 81091, Bruker instrument, Inc., 1981. M. R. Yemen, Ph.D. Dissertation, Michigan State University, 1982. J. Papaioannou, unpublished results, this laboratory. J. L. Dye and V. A. Nicely, J. Chem. Ed. 38, 43 (1971). R. S. Alger, "Electron Paramagnetic Resonance: Techniques and Applications", Interscience Publishers, New Ybrk, 1968. D. C. Mhttis, "The Theory of.thnetism", Harper & Row, New York, 1965. 0. Fussfi, S. Kauzlarich, J. L. Dye and B. K. Teo, J. Am. Chem. Soc. A8Z, 3727 (1985). O. Fussé, B. K. Teo and J. L. Dye, to be submitted to J. Am. Chem. Soc. R. K. Harris, "Nuclear Magnetic Resonance Spectro~ scopy", Pitman Books, Ltd., London, 1983. 67. 68. 69. 70. 71. 72. 73. 74. 75. 76. 77. 78. 79. 80. 81. 82. 83. 84. ~168- B. Lindman and S. Forsén, "NMR and the Periodic Table", (R. K. Harris and B. Mann, eds.), p. 129, Academic Press, New York, 1978. F. W. Wehrli, "Annual Reports on NMR Spectroscopy", (G.A. Webb, ed.), p. 126, Academic Press, New York, 1979. S. Forsén and B. Lindman, Meth. Biochem. Anal. 8Z, 270 (1981). K. J. Neurohr, T. Drakenberg, S. Forsén and H. Lilja, J. thh. Reson. 8A, 460 (1983). S. Khazaeli, J. L. Dye and A. I. Popov, Spectrochimica Acta 88A, 19 (1983). J. M. Ceraso and J. L. Dye, J. Chem. Phys. 8A, 1585 (1974). J. L. Dye, C. W. Andrews and J. M. Ceraso, J. Phys. Chem. Z8, 3076 (1975). A. Beckmann, K. D. B6klen and D. Elke, Z. Phys. gzg,173 (1975). , W. Lamb, Phys. Rev. 88, 817 (1941). N. C. Pyper and P.P. Edwards P. P. Edwards, S. C. Guy, D. M. Holton and W. McFarlane, J. Chem. Soc., Chem. Commun., 1185 (1981). P. P. Edwards, S. C. Guy, D. M. Holton, D. C. Johnson, J. Sienko, W. McFarlane and B. Wood, J. Phys. Chem. Z, 4362 (1983). m0: '0 P. Edwards, J. Phys. Chem. 88, 3772 (1984). R. C. Phillips, 8. Khazaeli and J. L. Dye, J. Phys. Chem. 88, 606 (1985). P. B. Smith, Ph.D. Dissertation, Michigan State University, 1978. J. L. Dye, in "Progress in Macrocyclic Chemistry", (R. M. Izatt and J. J. Christensen, eds.), Wiley- Interscience, New York, Vol. 1, p. 63, (1979). E. R. Andrew, W. S. Hinshaw and A. Jasinski, Chem. Phys. Lett. 88, 399 (1974). E. R. Andrew, Prog. NMR Spectrosc. 8, l (1971). 85. 86. 87. 88. 89. 90. 91. 92. 93. 94. 95. 96. 97. 98. 99. 100. 101. 102. 103. ~169- P. Mansfield, Prog. NMR Spectrosc. 8, 41 (1971). U. Haeberlen, Adv. Magn. Reson., Suppl. (1) (1976). M. Mehring, NMR Basic Principles & Prog. AA (1976). "NMR Spectrscopy in Solids”, (Proceedings of a Symposium), The Royal Society (London) (1981). A. C. Cunningham and S. M. Day, Phys. Rev. A88 (287), (1966). S. Ganapathy, S. Schramm and E. Oldfield, J. Chem. Phys. ZZ, 4360 (1982). K. Narita, J.~I. Umeda and H. Kusumoto, J. Chem. Phys. 85, 2719 (1966). E. Kundla, A. Samoson and E. Lippmaa, Chem. Phys. Lett. 88, 229 (1981). v.0. Mfiller, Annalen der Physik Z, 451 (1982). D. Freude, J. Haase, J. Klinowski, T. A. Carpenter and G. Ronikier, Chem. Phys. Lett. AA8, 365 (1985). A. Samoson, Chem. Phys. Lett. AA8, 29 (1985). H.~J. Behrens and B. Schnabel, Physics AAA8, 185 (1982). E. Oldfield and R. J. Kirkpatrick, Science 88Z, 1537 (1985). A. Ellaboudy, M. L. Tinkham, B. Van Eck, P. B. Smith and J. L. Dye, J. Phys. Chem. 88, 3852 (1984). J. L. Dye and A. Ellaboudy, Chem. Br. 88, 210 (1984). E. Oldfield, S. Schramm, M. D. Meadows, K. A. Smith, R. A. Kinsey and J. Ackerman, J2 Am. Chem. Soc. A88, 919 (1982). M. D. Meadows, K. A. Smith, R. A. Kinsey, T. M. Rothgeb, R. P. Sarjune and E. Oldfield, Proc. Nat. Acad. Sci. USA Z8, 1351 (1982). A. Ellaboudy and J. L. Dye, J. Magn. Reson., in press. P. P. Edwards and A. Ellaboudy, Nature, 8~Z, 242 (1985). -170- 104. M. L. Tinkham, A. Ellaboudy, P. B. Smith and J. L. Dye, J. Phys. Chem., in press. 105. J. H. VanVleck, Phys. Rev. Z8, 1168 (1948). TY MICHIGAN STATE UNIVERSI LIBRARIIES IIIIllfllilH‘IlfllllIlmlljllllililllllIllifllllHliHllHllI I 3193 3177 594 I