xx ‘ .. ..( ‘13:... . 11...}... \. ;..« L. 3...; (‘16.... ‘ \lilllllll\ll\lll\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\ This is to certify that the thesis entitled lVblecular Magnets from Stable Organic Free Radicals: An Ion—Binding Approach presented by Sei—Hum Jang has been accepted towards fulfillment of the requirements for Ph.D. degree in Chemistry QM filo Major prof or Date (O//717/q} 0-7639 MS U is an Affirmative Action/Equal Opportunity Institution _ LIBRARY Michigan State ' University PLACE IN RETURN BOX to remove this checkout from your record. TO AVOID FINES return on or before date due. DATE DUE DATE DUE DATE DUE "33.04.1225, L , I MSU Is An Affirmative Action/Equal Opportunity lnditution , , 7 , , Warn-9.1 Molecular Magnets from Stable Organic Free Radicals: An Ion-Binding Approach by Sei — Hum Jang A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Chemistry 1993 ABSTRACT Building Organic Magnets from Stable Organic Free Radicals: An Ion-Binding Approach by Sei - Hum Jang Structural studies of the propeller-shaped radical 1 and related triaryl-X species show pairs of binding sites comprising tripods of ether oxygens. The preorganized nucleophilic pockets have enough flexibility to accept a range of metal ion sizes. By coordination with small alkali metal cations the radicals may stack to make extended chains of interacting electron spins as shown below. In fact, mixtures of tris(2,6- dimethoxyphenyl)methyl 1 and LiBF4 form solids with weak ferromagnetic behavior and solids from 1 and CdC12 show clear antiferromagnetic behavior. nX' nX' nX' M = u 2 (RMR dimer) 3 (RW oligomer) x = BF4‘, BPh4’, r, 0104‘, SCN' NMR, pulsed EPR, UV—Vis, and electrochemical studies of radicals and diamagnetic analogs show behavior consistent with ion binding to form complexes. In addition, by varying substituents on the aryl rings of 1, it is possible to adjust redox and binding properties and thus the radical's electronic and structural nature. Such adjustments are significant tools for fine tuning radicals for effective ion binding and electron coupling. Extended oligomeric arrays are difficult to characterize, so model compounds that form well-defined molecular complexes were devised to separate and characterize the metal-radical, metal-radical-metal, and radical—metal-radical relationships. Because magnetism is a bulk property that arises from a three-dimensional structure, potential chain-coupling di- and tri-radical building blocks were devised by utilizing m-phenylene as a high spin coupling unit. Magnetic measurements with a Superconducting Quantum Interference Device (SQUID) magnetometer show that, unlike l-LiX, some of monomeric complexes show simple paramagnetism, with independent electron spins. Control over magnetic properties can be achieved by varying the metal cation, thus complex 1-LiX shows weak ferromagnetism while complex 1-CdC12 shows antiferromagnetic coupling. Electron Spin Echo Envelope Modulation (ESEEM) studies of complexes in the radical ionophoric system 28 in frozen matrices yield the number of metal ions bound and their distances to the radical center. Molecular mechanics and semiempirical molecular orbital (MO) calculations coincide remarkably well with the spectrosc0pic findings. Complexation studies in solution support these ideas while Ab Initio calculations on a linear H3C-Li+-CH3 model show that high-spin (ferromagnetic) electron coupling is favored for a wide range of radical-Li+ distances. Model Ab Initio MCSCF computations on the same system offer an explanation for high—spin complexation-induced magnetic coupling of radicals. Attempts to assemble and characterize a series of triarylmethyls and their metal ion complexes were made. Triaryl-X frameworks are new to the ion binding field, so the complexation abilities of 1 and its congeners were studied by UV-Vis, electrochemistry, NMR and EPR methods. ESR, SQUID, and X-ray studies on these systems were used to asses inter- electron communication. Besides addressing structural and theoretical issues of magnetic coupling, the requirements for the design of molecular solids with long-range structure induced by ion complexation were probed. To my Parents ACKNOWLEDGMENTS My sincere thanks are given to Professor James E. Ned Jackson whom I believe to be one of the great chemists. It was my honor to be his first student and is going to remain as a happy memory in my heart. Numerous discussions with him on my research and others made me realize the meaning of learning how to learn. I would like to thank Professor Bart E. Kahr, Professor John McCracken, and Hong-In Lee for their help in my research and Professor James L. Dye, Professor William Reusch, and Professor Daniel G. Nocera for serving on my committee. I am grateful to the National Science Foundation, the United States Air Force, and the Michigan State University Center for Fundamental Materials Research for financial support in the form of teaching and research assistantships. I would like to thank to the Department of Chemistry for its excellent research environment. I also owe special thanks to many friends in the Chemistry and Physics departments for their friendships. Last, I would like to thank my wife and family for love and patience throughout my study. TABLE OF CONTENTS Chapter page LIST OF TABLES ......................................................................... viii LIST OF FIGURES .......................................................................... x INTRODUCTION ......................................................................... l 1. Macromolecular Chemistry and Molecular Engineering ................... 4 2. Stable Organic Radicals ................................................................ 6 2.1. Tri- and Diphenylrnethyl Radicals Historical Background .................................................... 7 2.1.1. Gomberg's Triphenylrnethyl Radical ............................... 7 2.1.2. Schlenk, Chichibabin, and Thiele's Hydrocarbon ............... 9 2.1.3. Ballester's Stable Triaryhnethyl Radicals ......................... 11 2.1.4. Jackson and Kahr's Lithium Complex of Triarylrnethyl ..... 13 2.2. Nitrogen Centered Radicals ............................................ 15 2.2.1. Hydrazyl Radicals .......................................................... 15 2.2.2. Nitroxyl Radicals ........................................................... 17 2.3. Aroxyl Radicals ............................................................. 21 3. Magnetism ................................................................................... 23 3.1. Diamagnetism ............................................................... 23 3.2. Paramagnetism .............................................................. 24 3.3. Curie-Weiss Law ........................................................... 28 3.4. Pairwise Magnetic Exchange ........................................... 30 3.5. One-Dimensional or Linear Chain Systems ....................... 31 3.6. Alternating Linear Heisenberg Chain ............................... 35 4. Molecular Magnets ....................................................................... 37 4.1. Molecular Magnet models ............................................... 37 4.1.1. McConnell Model .......................................................... 37 4.1.2. Mataga Model ............................................................... 39 4.1.3. Ovchinnikov Model ....................................................... 41 4.1.4. Breslow Model .............................................................. 42 4.1.5. Torrance Model ............................................................ 44 4.1.6. Kahn's Proposal ............................................................ 44 V 4.2. Metal Based Molecular Magnets ...................................... 46 4.2.1. Single Atom Bridged Bimetals ........................................ 46 4.2.2. Molecule Bridged Bimetals ............................................. 48 4.2.3. Metal and Organic Radical Based Systems ........................ 49 4.2.4. Charge Transfer Complex Based Systems ......................... 54 4.3. Galvinoxyl and N itroxyl Radical Based systems ................ 55 4.3.1. Galvinoxyl Radicals ....................................................... 55 4.3.2. Nitroxyl Radicals ........................................................... 58 4.4. Altemant Hydrocarbon Based Polyradicals ....................... 63 4.4.1. Triplet Diradicals .......................................................... 66 4.4.2. Perchlorinated Triphenylmethyl Radicals ......................... 68 4.4.3. Polyphenyl Carbenes ...................................................... 72 5. Molecular Magnets by Macromolecular Chemistry .......................... 75 5.1. Long-range Magnetic Interactions in Organic Radicals ...... 77 5.2. Through Space Magnetic Interactions in Organic Radical Complexes .................................................................... 81 5.3. Design of Molecular Magnets by Ion-Binding ................... 83 RESULTS AND DISCUSSION .......................................................... 90 CHAPTER I. Triaryhnethyl Radicals ................................................. 90 1. Structure of Triaryhnethyl Radical ................................................ 90 1.1. Structure of T riaryhnethyl Radical .................................. 92 1.2. Structure of Triaryhnethyl Cation ................................... 95 1.3. Structure of Triaryhnethyl Borane .................................. 96 1.4. Hexachloro Triphenylmethyl Radical and Borane .............. 100 2. Substituent Effects ........................................................................ 102 2.1. Hexamethoxy T riphenylmethyl Radicals ........................... 104 2.2. The Cyclized Xanthenol ................................................. 120 2.3. Tetramethoxy Triphenylmethyl Radicals .......................... 121 2.4. Rotational Barriers ........................................................ 131 3. Ion-Binding .............................................................................. 137 3.1. Borane ......................................................................... 137 3,2. Double faced Paramagnetic Ionophore ............................. 141 3.3. Tetramethoxy Triphenyl amines ...................................... 153 3.4. Ammonium Complexes by Hydrogen bonding. ................. 160 3.5. Metal Ions ..................................................................... 165 vi a o n c i CCSILZAJ. E123 CHAPTER II. Magnetism of Triaryhnethyl Radical Complexes ............ 167 1. Calculations on Pairwise Interaction of Methyl Radicals ................... 167 2. Antiferromagnetic Interaction and Dimer of The Radical ................. 169 3. Magnetic Behavior of Radicals and Complexes ................................ 171 3.1. Diamagnetism of Methane and Metal Salts ........................ 173 3.2. Paramagnetism of Radicals and Radical Complexes ........... 175 3.3. Ferromagnetism of Complexes ........................................ 184 3.4. Antiferromagnetism of Radical Complexes ....................... 187 4. Design of Molecular Magnets by Other Ion-Binding ........................ 192 5. Introduction of Cross-Linkers ....................................................... 195 CONCLUSIONS AND SUGGESTIONS FOR FUTURE WORK ............ 199 Conclusions ...................................................................................... 199 Suggestions for future work .............................................................. 201 1. Aromatic Linked Biradicals .......................................................... 201 2. Magnetic Coupling through Hydrogen Bonds .................................. 202 3. Ion-Binding in Various N itroxyl Radicals ....................................... 204 EXPERIMENTAL ......................................................................... 208 1. General Procedures ...................................................................... 208 2. Solvents and Chemicals ................................................................. 209 3. Equipment and Procedures ............................................................ 209 CV Measurements .................................................................... 209 VT—EPR Measurements ............................................................ 210 SQUID Measurements .............................................................. 210 4. Synthesis ..................................................................................... 212 REFERENCES ............................................................................... 242 vii Ta LIST OF TABLES lble p age Aryl ring twists in tripod binding sites ....................................... 99 Crystal structure determination and refinement data for tris (2,6-dimethoxyphenyl)methyl 1, tris(2,6-dirnethoxyphenyl)methyl tetrafluoroborate 62, tris(2,6-dimethoxyphenyl)borane 63 ......... 99 Summary of electrochemical data for derivatives of tris(2,6-dimethoxyphenyl)methyl 1 ............................................ 105 Correlation of E1 /2 and o-values for derivatives of tris(2,6-dimethoxyphenyl)methyl 1 ............................................ 108 Summary of hyperfine constants for derivatives of tris(2,6-dimethoxyphenyl)methyl 1 ............................................ 116 Summary of electrochemical data for derivatives of phenyl-bis(2,6-dimethoxyphenyl)methyl 77 ................................ 123 Correlation of E1 /2 and o—values for derivatives of * phenyl—bis(2,6-dimethoxyphenyl)methyl 77 ................................ 125 Summary of hyperfine constants for derivatives of phenyl-bis(2,6-dimethoxypheny1)methyl 77 ................................ 126 Summary of VT-NMR results for tris(2,6-dimethoxy- phenyl)methyl ammonium tetrafluoroborate 72, tris(2,6- dimethoxy-3,5-dichlorophenyl)methyl ammonium tetrafluoroborate 75, tris(3,5-dichloro-2,6-dimethoxy phenyl)methane 76 .................................................................. 133 Complexation data from ESEEM for tris(2,6-di(2-methoxy ethoxy)phenyl)methyl 86*2MX ................................................. 1 50 viii Summary of electrochemical data for tris(2-(methoxy ethoxy)-6-methoxyphenyl)methyl 85 and tris(2,6—di- (methoxyethoxy)phenyl)methyl 86 ............................................ 151 . Summary of electrochemical data for bis(2,6-dimethoxyphenyl) -4-chlorophenyl amine 87, bis(2,6-dimethoxyphenyl) -3,5-dimethoxypheny1 amine 88, tris(3,5-dimethoxyphenyl) amine 89 ................................................................................ 155 . Crystal structure determination and refinement data for tris(2,6- dimethoxyphenyl)methyl ammonium tetrafluoroborate 72 and tris(3,5-dichloro-2,6-dimethoxyphenyl)methyl ammonium tetrafluoroborate 75 ................................................ 162 ix ll ;ure LIST OF FIGURES page Temperature dependence of the paramagnetic susceptibility ........ 25 The molar susceptibility for ferro- and antiferromagnets ............. 29 Bleaney Bowers susceptibilities compared with the Curie law ....... 31 Ising susceptibilities compared with Curie law ............................ 34 Ising "perpendicular" susceptibilities .......................................... 34 Heisenberg linear chain susceptibilities ...................................... 36 Illustration of the oligomeric chain of tris(2,6—dimethoxyphenyl)methyl 1 with M+ ions ....................... 87 MM2 optimized structure of an RMR dimer 2 calculated ............ 87 Stereo view of the X-ray structure of the radical tris(2,6-dimethoxyphenyl)methyl 1 ............................................ 94 Space-filling views of the binding site for tris(2,6- dimethoxyphenyl)methyl 1 and tris(2,6-dimethoxyphenyl) methyl tetrafluoroborate 62 ...................................................... 96 Ball & Stick representations of the binding site for tris(2,6-dimethoxyphenyl)methyl l and tris(2,6-dimethoxy phenyl)borane 63 .................................................................... 97 Stereo view of the packing diagram of tris(2,6-dimethoxy phenyl)borane 63 .................................................................... 98 MM2 calculated structure of tris(3,5-dichloro-2,6-dimethoxy- phenyl)borane 64 dimer with Li+ ............................................. 101 Cyclic voltammogram of tris(3,5~dichloro-2,6-dimethoxyphenyl) borane 64 in methylene chloride ............................................... 101 X 20. ' 22. 23. 24. 25. 26. 27. 28. 5. Cyclic Voltammogram of tris(2,6-dimethoxyphenyl) - \ methyl 1 in THF ...................................................................... 103 . AE1/2(V) vs 6 in THF for HMTP methyl radicals ....................... 108 '. EPR spectrum of tris(2,6-dimethoxyphenyl)methyl 1 in THF with simulations ........................................................... 117 . EPR spectrum of tris(3,5-dichloro-2,6-dimethoxyphenyl) methyl 65 in THF with simulations .118 . EPR spectrum of tris(4—chloro-2,6—dimethoxyphenyl) methyl 66 in THF with simulations 119 . The X-ray structure of cyclized tris(2,6-dimethoxy phenyl)methyl carbinol 120 AE1/2(V) vs 6 in THF for TMTP methyl radicals ....................... 124 EPR spectrum of phenyl-bis(2,6-dimethoxyphenyl) methyl 77 in THF with simulations ...... 127 EPR spectrum of phenyl-bis(2,6-dimethoxyphenyl) methyl 77 in toluene with simulations ........................................ 128 EPR spectrum of bis(2,6—dimethoxyphenyl)-4—chlorophenyl methyl 78 in THF with simulations ..... 130 EPR spectrum of bis(2,6-dimethoxyphenyl)-3,5-dimethoxy phenyl methyl 84 in THF with simulations ................................. 130 Stereo view of the X-ray structure of tris(Z,6-dimethoxyphenyl) methane 73 ............................................................................. 131 VT 300 MHz 1H NMR spectrum of tris(2,6—dimethoxy- 3,5—dichlorophenyl)methyl ammonium tetrafluoroborate 75 ........ 134 VT 300 MHz 1H NMR spectrum of tris(2,6-dimethoxy- 3,5—dichlorophenyl)methane 76 135 xi 31 3f 36 37 38 39 40. 41. UV-Vis spectra showing ion-binding in tris(2,6-dimethoxyphenyl)borane 63 139 Mass Spectrum of a dimer 2 of tris(2,6-dimethoxy- pheny)methyl 1-Na+ol ............................................................ 140 EPR spectrum of tris(2,6-di(methoxyethoxy)phenyl) methyl 86 with simulations ....................................................... 143 ESEEM spectrum of tris(2,6-di(methoxyethoxy)phenyl) methyl 86-2LiBF4 .................................................................. 145 ESEEM spectrum of tris(2,6—di(methoxyethoxy)phenyl) methyl 86°2NaBPh4 ............................................................... 147 Summary of ESEEM and MNDO results of tris(2,6— di(methoxyethoxy)phenyl)methyl 86-2LiBF4 ............................ 149 Stereo view of the MNDO calculated structure of tris(2,6-di(methoxyethoxy)phenyl) methyl 86-2LiBF4 ............... 149 1H NMR spectra for ion-binding of bis(2,6—dimethoxyphenyl)— 4-chlorophenyl amine 87 with Lil ............................................. 156 13C NMR spectra for ion—binding of bis(2,6—dimethoxyphenyl)— 4-chlorophenyl amine 87 with Lil.... ....157 1H NMR spectra for ion-binding of bis(2,6-dimethoxyphenyl)- 3,5-dimethoxyphenyl amine 88 with Lil .................................... 158 13C NMR spectra for ion-binding of bis(2,6-dimethoxyphenyl)- ,5-dimethoxyphenyl amine 88 with Lil .................................... 159 tereo views of the X—ray crystal structure of tris(2,6- imethoxyphenyl)methyl ammonium BF4 72 and tris(3,5- ichloro2,6—dimethoxyphenyl)methyl ammonium BF4 75 ............ 161 H NMR spectrum of complex of tris(2,6~dimethoxyphenyl) orane 63-NH4I in CD3CN ...................................................... 163 xii 42. 43. 1 44. 1 1H NMR spectrum of complex of tris(2,6-dimethoxy— 3,5—dichlorophenyl)borane 64-NH4I in CD3CN ......................... 164 Qualitative MO diagram for orbital interactions in MeLi+Me ....... 167 Curie—Weiss behavior of tris(2,6-dimethoxyphenyl)methyl 1 ...... 169 The X-Ray structure of head-to-tail peroxydimer 91 .................. 170 Plot of magnetization M vs H for tris(2,6-dimethoxyphenyl) methyl 1 and 1-LiBF4 ...172 Plot of magnetization M vs H for tris(2,6-dimethoxyphenyl) methane 73 and CdC12 ............................................................ 174 Plot of l/x vs T and xT vs T for tris(2,6—dimethoxyphenyl) methyl 1 ................................................................................ 176 Plot of M vs. H and l/X vs T for tris(2,6-di(methoxyethoxy) phenyl) methyl 86-2LiBF4 . ................................................... 179 Plot of ueff vs T for tris(2,6-dimethoxyphenyl)methyl 1-2LiI ....180 Plot of 1/x vs T for tris(2,6-dimethoxyphenyl)methyl 1-NH4I...181 Plot of Brillouin functions calculated for S=l/2 to 7/2 ................. 182 The saturation magnetization of the complex of tris(2,6-dimethoxyphenyl)methyl l-LiI at 1.8K .......................... 184 lot of magnetization M vs H for tris(2,6-dimethoxyphenyl)methyl ~LiBF4 and 1-ZnC12 .................. 186 lot of x vs T for tris(2,6-dirnethoxyphenyl) l-CdC12 ith Bleaney-Bowers fit ........................................................... 189 ltemating dimer and linear chain Heisenberg fit for is(2,6—dimethoxyphenyl) l-CdClz .......................................... 190 -ray powder patterns of 1 and l-CdClz .............................. 191 xiii _____A 0V1 an: int: sta‘ the are INTRODUCTION Organic materials science is in a stage of infancy compared to its inorganic sibling. The difficulties are associated mainly with the lack of control of structure in the organic solid state. Since the Nobel prize in 19871, interest in macromolecules and their self-assembly has increased ;lramatically. Attention has focused not only on the fundamental molecular 'ecognition processes of the macromolecules but also on the design of nolecular devices using various intermolecular interactions such as :lectrostatic, hydrogen bonding, van der Waals forces, etc. Organic materials chemistry requires the resources of traditional molecular hemistry combined with an understanding of non-covalent interactions so 3 to form macromolecular entities that possess structures as well defined as rose of molecules themselves. ne of the greatest challenges to materials science is the design of olecular magnetic materials. Simple paramagnetic organic compounds— :e radicals,2 triplet carbenes,3 biradicals4,5 —have been well studied er the last 30 years, and their magnetic characteristics can be adequately ilyzed in terms of current theory.6 Molecular magnets represent a damentally new class of materials in which a defined magnetic exchange :raction has been imposed on adjacent paramagnetic entities in the solid 6.7 In spite of significant growth in the field of molecular magnetism, mechanisms of colligative magnetic properties such as ferromagnetism incompletely understood.8 Interest in organic and molecular magnetism has grown so much in the last ten years that special ACS symposia on the subject were held in 1989 and 1992,9 yet to date there are only a few molecular magnets that clearly exhibit long—range ferromagnetic interactions.10 AF Coupler F Coupler . + i:> open shell singlet state + ‘ —-4— :9 open shell triplet state 11 an isolated atom or molecule, a pair of electron spins may be coupled to 'ield closed-shell singlet, open-shell singlet, or triplet states. In other lOI‘dS, we can relate the magnetic interaction between two organic radicals ) their chemical bonding interactions. Strong interactions between two {dicals will result in electrons pairing up in a closed shell singlet state (0— nd), while very weak interactions between two radicals will result in a ir of open shell doublet states (isolated paramagnets). We can view an tiferromagnetic coupling as a weak bonding interaction between radicals. extended odd-electron systems, the three modes above correspond to agnetic (closed-shell singlet), antiferromagnetic (open-shell singlet), ferromagnetic coupling (triplet), respectively.“ Several mechanisms 6 been discussed by which high-spin states of molecular materials may stabilized.12,13 However, there is still no predictive theory that can urately forecast the magnitude or even the sign of magnetic coupling in :eneric chemical system of known structure. Magneto-structural correlations are best described among structural homologs, and even then, experts do not always agree on the mechanisms of coupling. This thesis describes an approach to this problem that uses structurally well defined systems made of stable organic radicals. The principles developed in selective ion binding studies of polyether ligands14 suggest that the methoxy groups in tris(2,6-dimethoxy— phenyl)methyl radical can serve as binding sites for metal cations. Two radicals may "sandwich" a cation of appropriate size between them in a distorted octahedral pocket,15,16 fixing their relative orientation. The spatial relationships in such controlled radical aggregation can, in turn, dictate intermolecular electron coupling in "interrupted O'- bonds"———radical pairs or oligomers in which electron interactions are mediated by metal cations. Extended stacks formed by this complexation mechanism would ave at their cores a linear array of one-electron carbon-centered p- rbitals interacting through metal ions. In writing the first dissertation on this project in the Jackson group, I all have a rather extended introductory chapter to give a detailed picture f current molecular magnetism research, a brief overview of the theory of agnetism, and a historical review of stable organic radical chemistry. I all give the rationale behind our research, followed by results and scussion. Results and discussion will be divided into three parts: building uctural bases for the formation of complexes; magnetic studies on icals and complexes; and attempts at the introduction of linkers for tended magnetic structures. to en co en de sel im an. mo p05 of ent tha' latt C00 con 1. Macromolecular Chemistry and Molecular Engineering In contrast to molecular chemistry, which is predominantly based on the covalent bonding of atoms, supramolecular chemistry is based on intermolecular interactions. Supramolecular interactions are the foundation for highly specific biological processes, such as substrate binding by enzymes or receptors, formation of protein complexes, intercalation complexes of nucleic acids, and immunology. An exact knowledge of the energetic and stereochemical characteristics of these non-covalent, multiple intermolecular interactions within defined structural motifs should allow the design of artificial enzymes or receptors, which bind substrates strongly and selectively.17 There are many examples of such receptor or enzyme models including cyclodextrins,18 and cyclophanes based on bipheny119, terphenyl, nd triphenylmethyl20 frameworks, and cryptophanes.21 Molecular materials are characterized by being made up of discrete olecules. The structural properties of molecular materials offer many ossibilities to modulate the bulk electrical, magnetic, and optical properties the material by choosing appropriate molecules as the macromolecular tities. At the same time, it is a challenge to develop synthetic strategies at allow the control of the spatial distribution of the molecules in the ttice. The bulk properties of materials are always determined by operative interactions between the constituting molecules, which -nsequently must be assembled in the lattice in such a way as to maximize 3 bulk response.22 re 311 00 Beyond the basic problem of establishing correlations between structure and a given property, molecular materials appear promising for development of new characteristics through the combination and linking of their properties. The recognition of inter- and intramolecular interaction in macromolecules is the basis for the design of molecular materials.7 An organic crystal is perhaps the most precise example of molecular recognition, since molecular chemistry blends with intermolecular forces to propagate a periodic molecular array.23 The almost perfect alignment of molecules in an organic crystal usually results in highly regular physical and chemical properties of the molecules, which in turn justify efforts in :rystal engineering. Of course, phenomena such as hydrogen bonding and nolecular complexation have been known for decades. However, ideas :onceming molecular recognition have only been refined in recent years. The research on organic materials with nonlinear optical properties d the studies on organic conductors and superconductors are some of the ost active areas in the field. Much of the motivation for the increased search activity in conducting materials has been based on the potential for chnological applications. Conducting, light-weight, soluble, and ermoplastic materials are of great industrial and economic interest. ning the past ten years, there has been a very significant international search effort as a result of the search for higher Tc organic )erconductors, much of it stimulated by the discovery of the high Tc )per oxide ceramic superconductors.24a25 Til CC In: 3V 2. Stable Organic Radicals A free radical is an atom, molecule or complex which contains one or more unpaired electrons. Free radicals can be obtained in a variety of ways; those which can be isolated and exist as stable, pure substances may be prepared by conventional chemical methods but the more reactive radicals are produced by thermal or photochemical bond cleavage, decomposition, irradiation, mechanical degradation, or by electron transfer reactions. The majority of free radicals obtained by these methods are highly reactive and have lifetimes measurable in terms of micro— or milliseconds unless they are stabilized by trapping in some inert matrix or on the surface of a solid. The concept of stability in organic radicals depends on the chemical :ircumstances; for example, the methyl radical can be stabilized indefinitely an argon matrix at very low temperatures although it disappears eversibly on warming the matrix. Diphenylpicrylhydrazyl (DPPH) can ist permanently as a free radical in the solid state at room temperature; it n be crystallized, isolated and treated as a normal organic compound but 'hen exposed to other organic radicals it usually reacts rapidly to give amagnetic products. The word "stable" should only be used to describe a dical so persistent and so unreactive to air, moisture, etc., under ambient nditions that the pure radical can be handled and stored in the lab with no we precautions than would be used for the majority of commercially lilable organic chemicals.26,27 2.1. Tri- and Diphenylrnethyl Radicals: Historical Background 2.1.1. Gomberg's Triphenylmethyl Radical In 1900 Gomberg discovered triphenylmethyl, the first organic free radical. The historical significance of the discovery was reflected in "The History of Organic Chemistry in The United States, 1875—1955" by D. S. & A. T. Tarbell.28 " Without question, the single most important piece of work from the United States in the beginning of the century and the one which aroused the rnost interest abroad was the discovery by Moses Gomberg of structures which contained trivalent carbon. Gomberg was an example of the best rind of American success story. " " Coming to this country from Russia as child with his penniless father, he eventually worked his way through the niversity of Michigan at Ann Arbor and received a doctorate in chemistry ere in 1894." "During a year of study with Victor Meyer at Heidelberg in 97, he synthesized tetra-phenyl methane, and after his return to Michigan, 3 studied the preparation of the hexaphenylethane by the action of various etal salts and metals on triphenylmethyl chloride. He obtained, instead of e expected hexaphenylethane, a compound whose composition and emical properties corresponded to the peroxide." mberg suggested in his first paper that the highly reactive substance from reaction of silver metal with triphenylmethyl chloride contained alent carbon. Gomberg's interpretation of the triphenyl methyl was g against the classical structural theory of tetravalent carbon. From the same reaction under nitrogen, he isolated the first product and showed that it had the composition corresponding to hexaphenylethane by elemental analysis. However, the structure of the first product was not proven by Gomberg to be hexaphenylethane. The elegance and serendipity of Gomberg's experiments in 1900 have made the discovery familiar to many organic chemists. The significance of the claim of an isolable trivalent carbon molecule was immediately recognized, and leading chemists of the day entered the debate on the nature of Gomberg's hydrocarbon 4. QiHiQ Gomberg Dimer he hypothesis of an equilibrium between triphenylmethyl radicals and hexaphenylethane had been accepted as certainly as that of tetrahedral :arbon until the discovery by Lankarnp, Nauta, and MacLean in 1968 that hexaphenyl ethane" in fact had the unsymmetrical quinoid structure 5 vhich Jacobson had proposed for it in 1904. 29 o o —e 00 So far, Gomberg's hexaphenylethane has not been synthesized except the derivative hexakis(3,5-di—t—butyl-4-biphenyl) ethane by Winter in 78 and hexakis(3,5-di-t-butylphenyl) ethane by Mislow and Kahr in 1986.30 The realization that trivalent carbon could exist as a relatively stable structure made chemists more ready to accept free radicals as intermediates in reactions in solutions even when the individual radicals could not then be identified. 2.1.2. Schlenk, Chichibabin, and Thiele's Hydrocarbon On dehalogenating 1,3—bis(diphenylchloromethyl)benzene Schlenk and Brauns obtained a colorless product whose elemental analysis accorded with the molecular formula of Schlenk biradical 6. Investigation showed the biradical to be associated into a polymer.31 Ph Ph Ph Ph Schlenk Biradical 6 Sfimmermann produced the monomeric Schlenk biradical by electrochemical reduction of the corresponding dication with a rotating platinum electrode and assigned the triplet EPR spectrum with the zero- field splitting parameters of D =0.0079 cm'l, E S 0.0005 cm'1 for Am=1. The zero-field splitting parameters D and E can be determined directly from the triplet EPR spectrum; D and E values provide information on the distance between radical centers and the molecular symmetry of the triplet, respectively. The three levels of a triplet EPR spectrum (according to x, y, and z axis) are split even in the absence of an external magnetic field. This ero field splitting (zfs) arises from a dipolar coupling of the two spins that creates an internal magnetic field in the molecule, which splits the energy levels. Siimmermann could not detect the half field transition, a hallmark of a high spin state, from the spectrum of Schlenk biradical.32 The "half—field spectrum" corresponds to Am=2 transitions in a triplet EPR spectra. This transition is quantum-mechanically forbidden, but in most organic - diradicals, the zero field splitting relaxes the selection rule and the transition can be observed. It is a critical spectral feature because its presence unambiguously signals that one is observing a triplet state. Shortly after Gomberg's preparation of the triphenylmethyl radical in 1900, Chichibabin attempted to synthesize the analogous diradical 7.33 He obtained a blue-violet compound that reacted rapidly with oxygen, yielding polymeric peroxide. Early studies on Chichibabin's hydrocarbon 7 led to peculation as to whether it existed as a singlet, a triplet, or a mixture of the wo spin states.34 Numerous attempts to observe triplet spectra in solution ave failed; doublet spectra were found, however. Platz discussed existing xperimental inconsistencies and assigned a singlet ground state to hichibabin's hydrocarbon. Montgomery reported a crystal structure of hichibabin's hydrocarbon and concluded that it has a singlet ground state ith an unusually large amount of "diradical character" by comparing each nd length in the molecule with di-p-xylylene framework.35 He only corded a singlet EPR spectrum even with the single crystal of the adical used for the X-ray structure determination. 1 1 Ph Ph Ph Ph Chichibabin‘s Biradical 7 Thiele's Hydrocarbon 8 Thiele synthesized the first isolable derivative of p-xylylene 8 in 1904, after attempts to prepare the unsubstituted compound failed.36 Thiele‘s hydrocarbon 8 reacted with oxygen, but could be manipulated without elaborate precautions. The Thiele's hydrocarbon has been examined by a variety of physical methods, and there seems to be little doubt that the molecule has a singlet ground state. Montgomery reported a crystal structure of Thiele's hydrocarbon and concluded that it has a singlet ground state. 2.1.3. Ballester's Stable Triarylmethyl Radicals Since Gomberg's discovery of the first free radical a great number of stable organic radicals have been detected and isolated. Although carbon centered free radicals are highly reactive species, some of them show emarkable stability, the best example being Ballester's erchlorotriphenylmethyl radical series shown below.37 Some of these adicals have half-lives of decades in the air, and withstand typical radical eagents like nitric oxide, hydroquinone, quinone, and even highly ggressive chemical species like concentrated sulfuric, nitric acids, sodium ydroxide, or halogen, with little or no alteration. Also, they possess a markably high thermal stability, up to 300 °C. They are all completely 'sassociated in solution. 12 R Ballester‘s PMT radicals R R. R'. R"=C1. C6C15 R =H. Br. 1. CH3, cozn, C02CH3, co,NH4 These radicals have distorted D3 propeller structures (C2) in the solid state, and their high stabilities have been traced predominantly to the shielding of the trivalent carbon atom by their phenyl rings with a propellerlike :onformation and the six ortho chlorines. Thermal and chemical stabilities f free radicals are governed by two main factors: resonance and steric indrance.38 The inhibition of resonance and steric hindrance frequently ause high thermal stability and chemical inertness of the aromatic lorocarbons.39 The perchlorotriphenylrnethyl radical has twist angles of .3, 53.4, and 538° between the phenyl groups and the reference plane of e central carbon in the solid state. The effect of the absence of ortho or eta chlorines on the twist angles of the rings, and therefore on their bility, has been assessed by direct comparison of X-ray structures of rious derivatives of Ballester's radical by Veciana and coworkers. The ults were consistent with the steric shielding of chlorine atoms being the in reason for the observed stabilities of these radicals. —M‘ m ‘A 7"; V: ‘2 13 2.1.4. Jackson and Kahr's Lithium Complex of Triarylmethyl Tris(2,6—dimethoxyphenyl)methyl 1, originally synthesized by J. C. Martin, is a remarkably stable organic free radical.40 It can be synthesized efficiently using the procedure shown below. It is air-stable, presumably due to the fact that in its D3 propeller conformation the central trivalent carbon is protected from above and below by six ortho-methoxy groups similar to Ballester‘s radicals. Martin reported detailed EPR studies of the radical 1 and its 2D and 13C (central carbon) derivatives with the correlation of hyperfine coupling constants and the average twists of aryl rings of 47° in solution. Li 0/ E5 0 O/ MeOJLOMe ——> ————->~ HOC O O 3 \ / \ / O 0 11+ +C CI'Clz C O 3 O 3 \ \1 In 1985 Jackson and Kahr treated solutions of the radical 1 with solutions of lithium salts in order to study the possibility of complex formation between the radical 1 and metal cations. Precipitates formed upon mixing that requently gave unusual, intense ESR spectra. Some of the spectra were ery broad, others had considerable fine and hyperfine structure, and often —values deviated markedly from the free electron value.41 During the ollowing year they unsuccessfully tried to establish a structural basis for 44 14 the surprising magnetic behavior of their precipitates by growing single crystals from the precipitated solids. Results at that time suggested colligative magnetic properties as a possible explanation for their puzzling and highly variable ESR spectra. Their proposed model did not include any of the structural elements suggested by available strategies for achieving ferromagnetic coupling of spins between carbon centered radicals.42,43 The general line of research was set aside because, at the time of the initial discoveries, the principle investigators were graduate students whose chief responsibilities were unrelated to the present subject or the chemical system described. However, in 1989, Professor Jackson reexamined precipitates of 1-LiBF4 and l-LiClO4 with a Superconducting Quantum Interference Device (SQUID) magnetometer in the Department of Chemistry (Professor Dye's SQUID) at Michigan State University. While a polycrystalline sample of 1 behaved as a paramagnet, the susceptibility of the "salted" precipitates showed a ferromagnetic field— dependent hysteresis, signatures of colligative magnetic behavior. Surprised by these new results Professor Jackson revived investigations of organic magnetism with the aim of establishing a rational basis for the observed magnetic phenomena. We are now pursuing this research with the expectation that a deeper understanding of electron coupling between simple aramagnets will result. 15 2.2. Nitrogen Centered Radicals 2.2.1. Hydrazyl Radicals The discovery and early investigations of hydrazyl radicals were made mainly by Goldschmidt and coworkers during the period of 1920 to 1929.44 From his original observation that treatment of triphenylhydrazine with lead dioxide in ether gave a deep blue solution which rapidly changed to green and then a red—brown, he proposed that the labile blue intermediate was triphenylhydrazyl 9 and the final product hexaphenyltetrazene 10. Q QNN $0.14 2:30 (1,6) 9 10 e equilibrium between radical and dimer was further confirmed by the ependence of the blue color on the polarity and temperature of the solvent. oldschmidt also prepared and characterized various hydrazyl radicals cluding the remarkably stable, 2,2-diphenyl-1-picrylhydrazyl (DPPH); hown on the next page which shows no tendency to dimerize in the solid ate or in solution even at room temperature.45 Because of its applications, terest in hydrazyl chemistry has been focused on DPPH. It has been used nce 1950 as a radical scavenger in polymer chemistry and radiolysis and a lesser extent in synthetic organic chemistry.46 No2 DPPH Very early magnetic susceptibility measurements established that under zarious conditions DPPH was essentially 100% free radical.47 It has been ised as a reference radical frequently in connection with EPR spectroscopy )ecause of its availability and stability. The EPR spectrum of crystalline ;amples of DPPH was measured and found to consist of one narrow line tbout 1.7 gauss wide, but in a dilute benzene solution it shows a five line spectrum which was originally attributed to equal coupling of the unpaired :lection with the two nitrogen atoms, aN1=aN2= 8.9 gauss. More refined alysis of the spectrum revealed that the two nitrogens in DPPH were not quivalent and had coupling constants of 9.35 and 7.85 gauss. The larger of ese two constants was shown to be associated with the picryl nitrogen by 5N labeling studies.48 Apparently, for some reason, there are no studies n these radicals as base radicals for organic magnetic systems. They ould have significant potential for such applications in comparison with er nitro functionalized organic radicals. l7 Oxidation of the dihydrazine 11 with lead dioxide yields unstable pentaphenyl picryl mono hydrazyl (PPPMH) 12 which has hyperfine coupling constants similar to other meta-substituted picryl hydrazyls . Further oxidation gives the very unstable "pentaphenyl picryl dihydrazyl (PPPDH)" 13, which is considered to be diamagnetic below -30 °C from EPR and NMR measurements. Some workers claim to have obtained the ' EPR spectrum of the biradical form shown below.49 2.2. Nitroxyl Radicals Porphyrexide 14 was the first organic nitroxyi radical to be isolated .d characterized even though it was formulated as a "derivative of iadrivalent nitrogen.50 A number of diarylnitroxides 15 were bsequently prepared and characterized as radicals which do not dimerise the solid state or in solution (R=OMe). 51 18 r 9 NH R : ‘R HN R=H,OMe 14 15 2,2,6,6-tetramethylpiperidine-N-oxy1 (TEMPO) 16 has been studied extensively, and its Lewis basicity has been studied mainly by Drago.52 EPR spectra of TEMPO in solution have been reported to be solvent sensitive. Drago tried to explain the facts by the correlation of changes in :he hyperfine coupling constant on adduct formation with enthalpies of interaction and with changes in the infrared spectra of the hydrogen- aonding acid upon interaction. The nitrogen hyperfine coupling constant is found to increase with the enthalpy of adduct formation. Quite a few Iariations on this sterically hindered nitroxyl have been reported, mainly 'rom France and Russia including 4-hydroxy TEMPO (TEMPOL) 17.53 O' Q {)4 N OH TEMPO 1 6 TEMPOL 1 7 . K. Hoffmann reported a series of t—butyl nitroxides prepared by ectrochemical or sodium reduction of nitrobutane in glyme, with a oposed mechanism shown below, in 50% yield. The method was not neral for the preparation of other di-t-alkyl nitroxides, but, t—butyl-2,6- nethoxyphenyl nitroxide 18 is prepared by hydrolysis of the reaction 19 mixture obtained from 2,6-dimethoxyphenyl-lithium and nitrobutane in 17.8% yield. 0' t Bu / t-B ' \ ,O' t-B t-Bu—NOZ & t-Bu—N‘“ Na+ u N+ Na+ fl» u\N—O' \ / \O- B x O' t—Bu t‘ u Ar - \ O]L1+ H20 Ar\ R/ \O R’N—O. 0' \O ('3' #14 £er 0 l 18 Diphenylnitroxide 15 (R=H) is only stable in dilute solution. In :oncentrated solution or in the solid state it behaves like t- )utylphenylnitroxide 19, decomposing spontaneously. Thus, the increase in .npaired electron delocalization, which occurs as a result of replacing a t- utyl group by a phenyl group, merely increases the number of reactive ites in the molecule and not its stability. '0. ,0. ©le CCU 19 s in the t-alkyl aryl series, however, stable diarylnitroxides are produced len C—O dimerization is prevented by the presence of blocking groups in = para—positions, thus, di—para-anisyl nitroxide 20 which is believed to be i first organic radical studied by single crystal X—ray and 4,4'- 20 linitrodiphenyl nitroxide 21 are stable for months. In this respect the :hlorine substituted system is apparently inadequate since it decomposes 'apidly. Di-(p-tolyl)nitroxide (di—p-methyl) is sufficiently stable to be solated but decomposes within hours.54 0 (I) 1'“ ONO \0/0 00/ N02 N02 2 0 2 1 )rtho-substituents in diaryl nitroxide stabilize the radical probably by 'orcing the phenyl groups out of conjugation with the radical center, hereby reducing unpaired electron density at the para-positions.55 Of >articular interest is bis-(2,6-dimethoxyphenyl) nitroxide 22 shown below. t has a very interesting structure not only by the similarity to our riarylmethyl systems, but also for the natural extension of our rationale in esigning molecular magnets by ion-bindings which I will discuss later. \0 O/ \O 22 21 2.3. Aroxyl Radicals It has long been appreciated that aroxyl radicals are intermediates in many phenol oxidations, and interest in these radicals has been enhanced by the fact that they play a role in many biological oxidations. Since the discovery of 2,4,6-tri-t-butylphenoxyl radical 23, attention has been given largely to hindered phenoxyl radicals such as galvinoxyl 24, some of which have been isolated and are sufficiently stable to be used as reagents.56 O' t-Bu t—Bu t-Bu t—Bu O' I I O t-Bu / t-Bu t-Bu 2 3 Galvinoxyl 24 Of special interest are a number of stable aroxyl biradicals. Oxidation of the quinone methide with lead dioxide gives first the mono-radical 25, its EPR spectrum shows interaction of the unpaired electron with the four equivalent ring protons of the conjugated rings (aH=1.3 gauss). On further oxidation Yang's biradical 26 is obtained in which there is delocalization between the three rings giving seven lines from the six ring protons (aH=0.86 gauss).57 22 2 5 Yang's Biradical 2 6 Yang's biradical 26 is the best characterized stable biradical with high spin multiplicities among conjugated systems.58 23 3. Magnetism 3.1. Diamagnetism If a sample is placed in a magnetic field H, the observed field with a sample will generally differ from the free space value. If the observed field with a sample is reduced, the substance is said to be diamagnetic.59 The molar susceptibility of a diamagnetic material is negative, and rather small {-1 to -100 x 10-6 emu/mol). Diamagnetic susceptibilities do not depend on field strength and are independent of temperature. Since organic :ompounds with unpaired electrons also have a number of filled shells, they nave a diamagnetic contribution to their susceptibility. Diamagnetic susceptibilities of atoms in molecules are largely additive, and this provides 1 method for the estimation of the diamagnetic susceptibilities of ligand itoms and counter ions in a complex like 1-MX. \OI’ Li+ : -1.0 x 10-66emu/g atom C104 : -32.0 x 10 emu/g atom (Q g (3104‘ C25 :-6.0 x 10'6-2mu/g atom x 25 H27 : -2.93 x 10 emu/g atom x 27 O 06 : -4.61 x 10.6 emu/g atom x 6 / § -2.89 x 104 emu/mol he Pascal constants provide an empirical method for this procedure. One Ids the atomic susceptibility of each atom, as well as the constitutive rrection to take account of such factors as n-bonds as shown above for the mplex l-LiClO4. This procedure is only of moderate accuracy, and the lues given could change from compound to compound. Greater accuracy 24 can be obtained by the direct measurement of the susceptibility of a diamagnetic analogue of the paramagnetic compound. 3.2. Paramagnetism A paramagnet concentrates the lines of force provided by an applied - magnet and thereby moves into regions of higher field strength. This results in a measurable gain in weight in Gouy or Faraday balances. Paramagnetic susceptibility is generally independent of the field strength, but it is temperature dependent. To a first approximation at high temperature, the susceptibility x varies inversely with temperature, which is the Curie Law: :2 Zr C is called the Curie constant, and T is the absolute temperature. Since x '1 = T/C, a plot of x '1 vs. T is a convenient procedure for the determination of the Curie constant; note that the line goes through the origin for paramagnetic materials in Figure 1. Since the magnitude of magnetic susceptibility at room temperature is an inconvenient number, it is common among chemists to report the effective magnetic moment, ueff , which is defined as l l l _ 3k — ~ _ 2 2 pay =<—fi)2( 571.5") N]. k 5Ei being the energy level separation between the level i and the ground state j. This model provides a basis for a simple derivation of the Curie Law, which states that a magnetic susceptibility varies inversely with emperature. For an organic free radical (S=1/2) in zero field, the two evels ms =i'l/2 are degenerate, but split when a field Hz is applied. 'he magnetic moment of a radical in the level n is given as “n = 5Ian/5H = -msgilB 1e molar macroscopic magnetic moment M is obtained as the sum over agnetic moments weighted according to the Boltzmann factor and it can approximated as following: in fol Heft 27 M ~ N gZuBZHz/4kT Static molar magnetic susceptibility is defined as x mol = Mmol/I‘I which is in the form of the Curie law where the Curie constant can be expressed as following when S=1/2. C = Ngztth /4k = 0.125g25(s +1)cm3K/mo/ This is a special case of the more general spin-only formula, where Heffz = gZuBzS(S+l) is the square of the "magnetic moment". Ng2u32S(S +1) Ate/)2 z = \ = N 3kT 3kT ft is interesting to examine the behavior of M in the other limit, of very arge fields and very low temperatures, the magnetization becomes ndependent of field and temperature, and becomes the maximum or aturation magnetization Msat which the spin system can exhibit. This ituation corresponds to the complete alignment of magnetic dipoles by the )plied field. M50! = Ng/‘lBS 28 3.3. Curie-Weiss Law The Curie law is the magnetic analog of the ideal gas law, which is expressed in terms of the variables p, V, T. For magnetic systems, one uses H, M, T and the thermodynamic relations derived for a perfect gas can be translated to a magnetic system by replacing p by H and V by 1/M. There are many situations in which the Curie law is not strictly obeyed. One source of the deviations can be the presence of an energy level whose population changes significantly over the measured temperature interval (magnetic phase transition); another source is the magnetic interactions which can occur between paramagnetic centers. To the simplest approximation, this behavior is expressed by a small modification of the Curie law, to the Curie-Weiss law, where the correction term, 6, has the units of temperature. 120w) When 6 is negative in sign it is called antiferromagnetic; when 9 is positive, it is called ferromagnetic. The constant, 0, characteristic of any particular sample, is best evaluated when T>10K , as curvature of x-1 usually becomes apparent at smaller values of T. 29 800 0.1125 X (emu/moi) 0.075 400 l/x (moi/emu: 0.0375 O 50 100 150 200 250 300 Temperature (K) .b) XT(emu/mol) he =+10K 0.5 , ‘ ‘_ e§=-10K (D l i i i ( 40 so 120 160 2 )0 Temperature (K) Figure 2. a)The reciprocal susceptibility, x'l, extrapolated from the high temperature data; b) xT, as a function of temperature. These plots represent a system in which g=2, S=1/2, as well as ferromagnetically coupled ( 9 =+10K) and antiferromagnetically coupled ( 6 = -10K) systems. 3.4. Blea Bleai For r. by se at J/i susce This zero ‘ positi the C 10w V1 2 whi mOIe . C0llplc the sa: bIOad’ 30 3.4. Pairwise Magnetic Exchange The magnetic susceptibility per mole of dimers was calculated by Bleaney and Bowers via the application of Van Vleck‘s equation (the Bleaney—Bowers equation). _ (2Ng2/132 /kT) 3 + exp(—2] / kT) For negative J (antiferromagnetic), x has a maximum. This may be found by setting dx /dT=0. With the definitions used here, the maximum occurs at J/kTm ~ -4/5. For ~J/k <> Tm)(at high temperature), the susceptibility follows a Curie—Weiss law, x = 3/[4(T- 0 )] with 6 = J/2k. This is a specific instance of the more general connection between a non- zero Weiss constant and the presence of exchange interaction. When J is positive, the susceptibilities calculated according to the above equation and the Curie law for spin S=1/2 do not differ greatly until temperatures very low with respect to 2.] /k are achieved; this is because of the extra factor of 2 which enters when comparing a mole of uncoupled S=1/2 spins with a mole of dimers. he susceptibilities for a pair of S=1/2 radicals antiferromagnetically oupled are discribed in Figure 3. While the low temperature behavior is he same, the temperature of maximum x increases with decreasing 0. A road, featureless peak is observed because of the continuous population ecrease of the various levels as temperature decreases. 311 (hassi dhner thh) behav Then: descn SySlen than“ 31 0.04 0.035 0.025 X (emu/mmol) 0.02 0.015 40 Temperature (K) Figure 3. The Bleaney Bowers susceptibilities compared with the Curie law; plots for J=0, -2.5, —5, -7.5, —10, -l2.5 cm'l. 3.5. One-Dimensional or Linear Chain Systems Much of what has been covered so far could come under the classification of short-range order. The physical meaning implied by low dimensional magnetism is that radicals are assumed here to interact only with their nearest neighbors in a particular spatial sense. This magnetic behavior follows directly from the structure of the various compounds. ere are good theories and extensive experimental data available which escribe the thermodynamic properties of one-dimensional magnetic ystems, at least for S=1/2; for systems of metal ions linked into uniform hains. In order to discuss exchange at the atomic level, one must intrt Hart This mod whet The I of th. Situai the I Heisc pract. There for a Chara in 1D exten S)lSter Since Calcu] Pledic ‘3' ’ A :‘z—V-mé—‘v' 32 introduce quantum mechanical ideas and we need to use the following Hamiltonian. 212—2st,- cs]- This is an isotropic Hamiltonian and is often referred to as the Heisenberg model. A very anisotropic expression, called the Ising model is obtained when 7 is set equal to zero in the following equation. H = —2JZ[y(51,52, + 31,52 , ) + 51,522] The Heisenberg model, which corresponds to using all the spin-components of the vectors 81 and 82, is obtained when 7:1. This appears as an artificial situation, and yet the Ising model is important, in part, because solutions of the Ising Hamiltonian are far more readily obtained than those of the Heisenberg Hamiltonian. Both of these Hamiltonians are restricted in their practical application to nearest—neighbor interactions. There are no exact or closed-form solutions for the Heisenberg model, even for a S=1/2 one—dimensional system. But, calculations are available which characterize Heisenberg behavior to a high degree of accuracy, particularly in 1D systems. Furthermore, one-dimensional short-range order effects are extended over a much larger region in temperature for the Heisenberg systems than for the Ising systems. Since the Heisenberg model is an isotropic one, the susceptibility is calculated to be isotropic. A broad maximum in the susceptibility is predicted for a linear chain antiferromagnet. For an infinite linear chain, 33 Bonner and Fisher calculate that the susceptibility maximum will have a value xmax/(N gZuBZ/IJI) ~ 0.07346 at the temperature kTmax/IJI ~ 1.282. The best example of an antiferromagnetic Heisenberg chain compound is [N(CH3)4]MnCl3, TMMC. The crystal consists of chains of J =5/2 manganese atoms bridged by three chloride ions. The antiferromagnetic 71D behavior in this nearly ideal Heisenberg system has been reviewed.60 The zero—field susceptibilities have been derived by Fisher for the S=l/2 Ising chain. Na .211 2 95,), = —%iu / 2kT)exp(J / kT), Ng 21132 2 ape, = ——p"l——[tanh(J/2kT)+(J /2kT)sech (J/2kT)]. Many susceptibility measurements are made on powdered paramagnetic samples, so that only the average susceptibility, , is obtained. The average susceptibility is defined as < Z >= (Xpar + 2;rper) 3 e meaning of the symbols "parallel" and ”perpendicular" refer to the xtemal magnetic field direction with respect to the direction of spin— uantization or alignment within the chains, rather than to the chemical or tructural arrangement of the chains. The zero-field susceptibilities for xpaI d xper are plotted in Figures 4 and 5. 34 0.05 004 ' : J=0 . J=-5 003 ' X (emu/mol) ' J='7'5 I J=-10 002 F : J=-12.5 001 h '+ +_+ . 3" , 0 llllllJlllllllllll O 20 40 60 80 100 Temperature (K) Figure 4 The Ising "parallel" susceptibilities compared with Curie law; plots for J: -2.5, -5, -7.5, -10, -12.5 cm‘l. =~20 )- \- X (emu/moi) 0.08 ,II'III[ll11I ll11‘lirll l I LPAJJJLLL o l. 1' f 20 4o 60 80 100 Temperature (K) 0 Figure 5 The Ising "perpendicular" susceptibilities for J: -20, ~30, -40, -50 cm-l. 44 inl alt int 6X fer SUE 0n alt: alt: the (let 35 3.6. Alternating Linear Heisenberg Chain The models discussed so far are uniform linear chain systems. The intrachain exchange constant has been assumed not to vary with position along the chain, and it is assumed that there is no interchain interaction. H = —212[s,-_1 . s,- + as, . [+1] .- An alternating chain is defined by letting 0t be less than one. For a Heisenberg alternating linear chain, -2] is then the exchange interaction between a spin and one of its nearest neighbors, and -2(XJ is the exchange constant between the same spin and the other nearest neighbor in the chain. When 0t=0, the model reduces to the dimer model with pairwise interactions (the Bleaney—Bowers model). The alternating linear chain antiferromagnet with S=1/2 has been studied in detail both theoretically and experimentally.61 In 1955 Oguchi developed a general theory of ferromagnetism and antiferromagnetism, based on the Heisenberg model of two—spin clusters. Ohya-Nishiguchi later extended this theory to the susceptibility of interacting spin-pair systems. One of the important susceptibility results for an antiferromagnetic alternating Heisenberg linear chain is a broad maximum for all values of the alternation parameter or. The susceptibility curves decrease exponentially as the temperature approaches zero.62 The zero—field susceptibilities have been derived by Fisher for the S=l/2 linear Heisenberg chain where u=(coth K- 1/K) with K=3J/2kT. Each susce] derive 4' is I SUSCeI chain 36 1 h 1— _23Ns#2 ”(“5317 T) /(23/=§§=-< 11%in N N N \ I \ O' O’ 0 Ullman's biradical n = l or 20 5 6 5 7 Very recently Ulhnan's biradical 56 and its derivatives have been studied by Rey. He reported X-ray structures and magnetic susceptibility measurements showing antiferromagnetic intramolecular interaction for the Ullman’s biradical. The singlet-triplet energy gap estimated was as -311 cm- 1 via the Bleaney-Bowers equation. Lahti reported computational modeling studies of various n-conjugated polyradicals using the AMI semiempirical molecular orbital method with configuration altemant hyd: polyradical n variety of cor groups as in Nitrene.1 19 5.2. Thro Complexes A ser constructed nitroxyl, an radicals as; use of Spin SuIIlmarize: complexes interactions E100mplex binding pr eempiex wr below, If fl centers, tilt sPeeifica1 ferr(magn: 81 configuration interactions which give theoretical basis for the heteroatom altemant hydrocarbons. He has been interested in synthesis of n-conjugated polyradical model systems based upon coupling of phenoxyl radicals by a variety of connecting spacer groups, such as heteroatoms and olefinic spacer groups as in the general model of Phenoxy-X-Phenoxy and Nitrene-X- Nitrene. 1 19 5.2. Through Space Magnetic Interactions in Organic Radical Complexes A series of spin labeled crown ethers (lS-crown-S) have been constructed by Mukai using stable organic radicals such as galvinoxyl, nitroxyl, and verdazyl (58, 59, 60, and 61) with the goal of using these radicals as probes for biological systems.120 There are good reviews on the use of spin labeled organic molecules for a biological probes and I will summarize mainly the systems with magnetic interactions. The alkali metal complexes of these radical crown ethers show interesting magnetic interactions in solution EPR .121 When the spin labeled crown ethers form a complex with alkali metal salts in solution according to the general ion- binding properties of crown ethers, lS-crown—S can only form a 1:1 complex with sodium salts but a 2:1 complex with potassium salts; as shown below. If there are significant magnetic interactions between the two radical centers, the EPR spectra of 2:1 complexes would reflect the interactions; specifically, a triplet spectrum would result when they are ferromagnetically coupled. warm Triplet spec galvinoxyl observed in system whit instead of p characterist: verdazyl rar triplet sper complexes systems wi potential it great exten \ 0'-N I H Triplet spectra have been reported for 2:1 complexes of nitroxyl 58, galvinoxyl 59, and phenoxyl radicals 60. Half field transitions were observed in all complexes except for the verdazyl labeled 15-crown-5 system which was reported as triplet coupled when sodium salts were used instead of potassium salts which is in contrast to the general ion-binding characteristics of lS-crown-S. The question remains whether the triphenyl verdazyl radical 61 itself forms a 2:1 complex with sodium cation to give a triplet spectrum. Furthermore, there is the possibility of forming complexes with different conformations (syn or anti) even in the above systems with triplet EPR spectra. Finally, the role of the oxy radicals as potential ion-binding sites for alkali metals has not been explored to any great extent. 00—b0 t—Bu w on ob&oo O 5 8 O 003 t-Bu 5 9 O 003 K/Ov’ K/Od O' t-Bu 5.3. Desigr The t1 traditionally ' it has been r ions in bimet interaction, overlap inte‘ orbitals. More recent such analyse Species, witl earlier. Int SUperexchar were made metal cente 01 tt-electro ttflk betwee energies be extended E btlnetaflic r twe‘eiectrc of the calc observe in 83 5.3. Design of Molecular Magnets by Ion-Binding The theoretical interpretation of superexchange interactions has traditionally been based on ideas developed for infinite solid lattices. Since it has been realized empirically that the bridging atoms between the metal ions in bimetallic systems determine the sign and magnitude of the exchange interaction, these qualitative treatments focus on the various types of - overlap interactions between the ligand atomic orbitals and the metal d orbitals. More recently there have been theoretical treatments which seek to extend such analyses to the cases involving molecular, rather than atomic bridging species, with special interest in molecular bimetallic complexes as discussed earlier. In this context a broader theoretical framework for the analysis of superexchange interactions has been proposed by Hoffmann.122 Attempts were made to show a connection between antiferromagnetically coupled metal centers and the phenomenon of through bond coupling of lone pairs or rt-electron systems in organic molecules. 123 Hoffmann established the link between antiferromagnetic exchange interactions and the difference in energies between degenerate MO's using the orbital energies obtained from extended Hiickel calculations (the simplest all valence-electron model) in bimetallic complexes. Although these calculations do not explicitly include two-electron interactions in organic radicals, he suggested that the behavior of the calculated orbital energies are expected to reflect what one would observe in more sophisticated calculations. The effect r exchange in geometricalr bond angle r same symme the metal orl earlier in bir Photoelectrr pair and re. interactions complexes affects the these Splitt metal-ligan respond to 84 The effect of the metal-single-atom-ligand-metal bridge angle on the exchange interaction has been studied extensively by Kahn.124 Such geometrical considerations lead one to expect a large AF coupling for a 180° bond angle when the metal orbitals can interact with a ligand orbital of the same symmetry, and a ferromagnetic coupling for a 90° bond angle when the metal orbitals are interacting through orthogonal ligand orbitals, as seen earlier in bimetallic systems. Photoelectron spectroscopy has provided evidence for the splitting of lone pair and n-levels as a consequence of through-space or through-bond interactions such as shown below. The bridging groups in dimetal complexes provide orbitals of a certain symmetry type, and this in turn affects the splitting of the metal orbitals. The interaction which leads to these splittings has a strong conformational dependence (180o angle for metal-ligand-metal) and metal centers coordinated to such systems should respond to the energy splitting by showing sizable magnetic coupling. _.1_N/ \N_1_ _.L_N// \\N_1_ | H | | \.=/ | Along the sa possible mag the electron between org substituents i As (1 Organic free carbon, the bulk of the ion binding can serve dtlnethoxy Size betWe Orientation The Spatia dietate int 85 Along the same line we might expect the systems shown below to have a possible magnetic coupling via bridging diamagnetic metals. Furthermore, the electron density on the metal may affect the degree of interaction between organic radicals as we have seen in reviewing the effect of substituents on bridging ligands in dimetal complexes. As discussed earlier, tris(2,6-dimethoxyphenyl)methyl 1 is a stable organic free radical and in its D3 propeller conformation around the central carbon, the unpaired electron is protected from above and below by the bulk of the six methoxy groups.125 The principles developed in selective ion binding studies of polyether ligands suggest that these methoxy groups can serve as binding sites for metal cations. Two of the tris(2,6- dimethoxyphenyl)methyl radicals can "sandwich" a cation of appropriate size between them in a distorted octahedral pocket, fixing their relative orientation. The spatial relationships in such controlled radical aggregation can, in turn, dictate intermolecular electron coupling126,127 in "interrupted 0- bonds", radical pairs metal cation would have orbitals inter with Kahn's "The orbitai easier to ac respects the up-down ty] space intera weak excep each other.‘ We are wo higher olig enforced b- might acco cations to them para eXelusivel WOIklng n 86 radical pairs or oligomers in which electron interactions are mediated by metal cations. Extended stacks formed by this complexation mechanism would have at their cores a linear array of one-electron carbon-centered p- orbitals interacting through metal ions.128 This idea happens to coincide with Kahn's recent suggestions for designing molecular magnets.129 "The orbital approaches based on the spin polarization effect might be easier to achieve. This approach is particularly attractive. Indeed, it respects the strong tendency of nature to favor local spin interactions of the up-down type." " To conclude, we would also like to stress that the through- space interactions on which we have focused in this Account are generally weak except when p atomic orbitals belonging to adjacent molecules point to each other." We are working to exert control over electron coupling in radical pairs or higher oligomers designed so that electron interactions are mediated and enforced by metal cations with varying electron densities. Various metals might accomplish this task, from diamagnetic alkali metal and alkaline earth cations to those of transition metals in various oxidation states, some of them paramagnets in their own right. Our efforts have focused almost exclusively on diamagnetic salts. The following scheme illustrates our working model: X=E Figur assoc Figr M01 radii Figr eltgomers Stttletures trtPOd bir. radical dir nX' O O mung-mm! nX' M = Li, Na etc. 2 (RMR dimer) 3 (RMn oligomer) X = BF4‘. BPh4', 1', C104‘, SCN' Figure 7. Schematic illustration of the dimeric and oligomeric association of tris(2,6-dimethoxyphenylmethyl) 1 with M+ ions. Figure 8. Optimized structure of an RMR dimer 2 calculated by the Molecular Mechanics methods in PCModel. Here the individual radicals adopt a homochiral, staggered conformation. Figure 7 portrays these "interrupted o-bonds" in radical pairs 2 or oligomers 3. CPK models, molecular mechanics calculations, and X-ray structures (vide infra) of related species support the metal ion-radical ether tripod binding. Figure 8 displays an illustration of a radical-metal ion- radical dimeric such as 2 as calculated by molecular mechanics. The a magnetic p: paramagneti straightforv rings and ir paramagnet stoichiomet 2:1, or 1:2 examples c molecule r possibility Thee Structure c electrons 1 mechanic: electron. Of structu Organic 1 aPillicatio tor develr layer a p model sy; gaVe [he 88 The above strategy targets rationally engineered materials whose magnetic properties result from collective interactions between simple paramagnetic centers. The basic structure of radical 1 allows for a range of straightforward modifications at the meta and para-positions of the aryl rings and in the group bound to the ether oxygen. By careful assembly of paramagnetic radical ligands, metal ions, counterions, and their respective stoichiometries, a variety of novel complexes maybe accessible. Simple 1:1, 2:1, or 1:2 complexes between metal ions and 1 in solution are monomeric examples of the desired types. Ultimately, extended chains in which each molecule of 1 is coordinated to two metal ions and vice versa offer the possibility of a designed molecular solid with unique magnetic properties. These would be fundamentally new materials, whose composition and structure can be easily to vary their bulk properties. With their unpaired electrons residing on light atom centers, organic free radicals are quantum mechanically simple with S=l/2 and g values close to that of the free electron. Our work may therefore shed light on the fundamental interplay of structure and magnetic character in materials.130 Since the problem of organic magnetism is essentially a problem in electronic structure, application of ab initio and semiempirical computational methods is essential for developing a detailed understanding of the electronic interactions that favor a positive exchange integral. Preliminary ab initio calculations on a model system CH3Li+CH3 have been conducted by Professor Jackson which gave the triplet as a ground state. The X-ray s in building ligand assoc Visible spec radical 1 W] are though interaction synthesized We have pu of our syst SQUID me markedly susceptibii aniferroma Detailed r Single Cry: Structure 2 robust ma 1 Will pm: the Stmc Propertiei terms of interactio at introdr coildllsk 89 The X-ray structures of related monomers provided structural information in building ligands for better ion-binding. Determination of the metal- ligand association processes in solutions are evidenced by NMR, EPR, UV- Visible spectrophotometry and cyclic voltammetry. The redox properties of radical 1 which can be adjusted by varying substituents of the aryl rings, are thought to be important in superexchange-mediated magnetic interactions. Thus, various substituted-triarylmethyl radicals were synthesized and studied electrochemically to determine the redox properties. We have pursued SQUID and EPR studies to examine the magnetic behavior of our systems as a function of temperature and magnetic field strength. SQUID measurements on various complexes of triarylrnethyl radicals show markedly nonlinear Curie-Weiss behavior and high paramagnetic susceptibilities, while polycrystalline powders of l-CdClz shows aniferromagnetic behavior. Detailed magnetic characterization of well-defined materials, ideally as single crystals, will enhance our understanding of the relationships between structure and odd-electron coupling. We hope these efforts will lead to a robust materials with controllable magnetic properties. I will present the results and discussion in the following order: i) building a the structural basis for complex formation; ii) studies on electronic properties of substituted radicals; iii) evidence for complex formation in terms of ion—bindings; iv) theoretical basis for pairwise magnetic interactions; v) magnetic behavior of radicals and complexes; vi) attempts at introduce magnetic cross couplers in building extended structures; vii) conclusions and suggestions for future studies. Illiuiilu 1” CHAPTER 1. Struct Unde is critical ti materials (1 properties. prediction analyzing ‘ determiner molecular The effect triphenyln estimated from HP] carbon. T at the 1: MeCOHne for all a gaussBl 90 RESULTS AND DISCUSSION CHAPTER I. Triarylmethyl Radicals 1. Structure of Triarylmethyl Radical Understanding the structures of organic compounds in the solid state is critical to the engineering of physical properties. The search for organic materials demands understanding of the forces that determine structure and properties. Such understanding should lead to improved methods for the prediction and design of organic crystal structures. The major challenge in analyzing the structure of molecular organic solids is that the structures are determined by summation of the many contributing weak intra and inter molecular forces. The effects of ortho or meta substituents on the twist angles of the rings in triphenylmethyl radicals, and therefore on their stability, have been usually estimated by comparison of the experimental hyperfine coupling constants from EPR spectra with theoretically calculated Spin densities on para carbon. The comparison can be achieved by the calculation of spin density at the para carbon from the hyperfine coupling constants using McConnell's equation, where Q is a factor that is approximately constant for all aromatic hydrocarbons and has an empirical value of -22.5 gauss.131 The twist at between the where [3 is t aromatic rir rings and t densities ar from hyper only 1H hy A systema electronic X-ray stru while crys ions have Of the ch cempared by Veciar To tris(2,6_d trustrater 02 toy crystal g settletura dttfractii SttttCturz 91 The twist angle is introduced by using B cos 0 for the resonance integral between the central carbon atom and the adjacent carbon atom in each ring, where B is the resonance integral between neighboring carbon atoms in in aromatic ring and 0 is the angle between the plane of one of the phenyl rings and the reference plane of the central carbon and its bonds. Spin densities are then obtained as functions of 0. But, these calculated angles from hyperfine coupling constants are not reliable enough, at least when only 1H hyperfine coupling constants are compared. A systematic study of the influence of substitution on geometries and electronic structures of chlorinated triphenylmethyl radicals based on their X-ray structures has been started by Veciana and coworkers in 1987132 while crystal structure determinations of many triphenylmethyl carbonium ions have been reported since 1965.133 The hyperfine coupling constants of the chlorinated triphenylmethyl radicals have been calculated and compared with those obtained in isotropic solution using INDO calculations by Veciana and coworkers. To date we have not obtained single crystals of metal complexes of tris(2,6-dimethoxyphenyl)methyl radical 1. Our efforts have in part been frustrated by the fact that Li+ catalyzes the oxidization of the radicals by 02 to yield the cation and the persistent cation in solution hampers the crystal growth process which I will discuss later. Consequently, our structural studies have focused on characterizations of the binding site via diffraction studies of the uncomplexed monomeric species and some close structural analogues and pulsed EPR studies. 1.1. Strut Tris(2 is a remark traced main rings whicl six ortho r coupling Cr rings in t comparing triphenylr nitropheny Martin est dimethox triphenyln 1. He p Substanti: SUbstitutir delocaliza the pmpr methoxy We were SOiution j allgles 01 “Sing Var 92 1.1. Structure of Triarylmethyl Radical Tris(2,6-dimethoxyphenyl)methyl 1, originally reported by Martin, is a remarkably stable organic free radical. The high stability has been traced mainly to the shielding of the trivalent carbon atom by the phenyl rings which are strongly twisted in a propellerlike conformation and the six ortho methoxy groups. Martin correlated observed EPR hyperfine coupling constants of the radical 1 with the twist angles of the aromatic rings in tris(2,6~dimethoxyphenyl)methyl radical in solution. By comparing the calculated twist angle of 30° for Gomberg's triphenylmethyl (aHpara=2.77), the X-ray structure of tris(4- nitrophenyl)methyl (0=30°), and a Hiickel Molecular Orbital calculation, Martin estimated the increase in angle of twist to be 17° for tris(2,6- dimethoxy-phenyl)methyl (aHpara=2.26) over that of Gomberg's triphenylmethyl resulting in an estimated twist angle of 47° for the radical 1. He pointed out that the reduction in apara from 2.77 to 2.26 is substantially larger than would be expected for simple methoxy substitutions in triphenyhnethyl radicals and that the decrease in spin delocalization should therefore be attributed to the increase in the pitch of the propeller conformation of the radical 1 induced by the six ortho methoxy groups. We were able to directly compare the solution and solid structures by using solution EPR and the X-ray structure of the radical 1. The estimated twist angles of the radical 1 in solution were nearly identical to Martin's values using various experimental conditions and solvents such as THF or diethyl ether (solut initially by Michigan S were obtair radicals adr twisted out two rings a the way I pathwayt unpreceder triaryhnetl Increased SUbstituen temPeratu: Sterically , tutti mini kctti/mol t We belies stable so] 93 ether (solution EPR). The X-ray structure of radical 1 was determined initially by Professor Kahr at Purdue University and repeated here at Michigan State University later. Brick-red monoclinic prisms of radical 1 were obtained by fast evaporation of an ether solution under argon. The radicals adopt an unusual conformation in the solid state. One aryl ring is twisted out of the central methyl carbon plane by only 12° while the other two rings are twisted by 61°. This structure represents a point well along the way to the transition state for the two-ring flip racemization pathway134 and its large deviation from the D 3 ground state is unprecedented for triaryl-X propellers, especially in a perorthosubstituted triarylrnethyl radical. Increased racemization barriers are consistent with large ortho- substituents; perchlorotriphenylrnethyl does not racemize on the room temperature time scale.135 However, methoxy groups are substantially less sterically demanding than chlorine atoms, and AMl calculations on 1 at the fully minimized and X-ray twist angles show an energy cost of less than 3 kcal/mol for this distortion. We believe that the X-ray structure of the radical 1 represent a kinetically stable solid state structure which may be far from the relaxed solution conformation. Because of the dissymmetric twists of the rings, the ether tripod bindi radical. The 1 average dis But, as note of energies Figr Small moi or Other ] results in the chara been con Removal radical. Pr the Li+ methoxy 94 tripod binding site is poorly represented by the crystal structure of the radical. The crystal structure of the radical l is shown in Figure 9. The average distance between the oxygen atoms of a tripod pocket is 3.7 A. But, as noted above, this distance can vary substantially over a small range of energies as a function of the three aryl rings' rotations. Figure 9. Stereo view of the X-ray structure of the radical 1. In pure solution, radical 1 is air stable; this suggests that even such a small molecule as 02 is sterically excluded. However, addition of LiBF4 or other lithium salts to air-saturated THF or ether solutions of radical 1 results in nearly instantaneous oxidation of the red radical solution to give the characteristic purple color of the triarylrnethyl cation. This process has been confirmed by Jackson and Kahr to be inhibited by excess l2-crown-4. Removal of air and reduction with acidic Cr(II) salts regenerates the radical. Professor Jackson proposed several possible interpretations. First, the Li+ ion may bind, as shown in the Scheme below, to the tripod of methoxyl oxygens in radical l, inducing a conformational change which opens up a binding site 02 itself, : molecule. i simply acti‘ 02 radical: Asr conceminp the sensiti 1‘2. Str A r cryStal st deviates PrOpefler and 3101 (MMpz) 95 opens up access for 02 to the radical center. Alternatively the fourth binding site on the (normally tetrahedrally coordinated) Li+ ion may bind 02 itself, serving as a bridge between the radical center and the 02 molecule. Finally, the presence of the Li't' ion in an organic solvent may simply activate electron transfer (ET) by strongly stabilizing the product 02 radical anion via ion pairing. As mentioned above, our attempts to obtain structural information concerning complexes of radical 1 with metal salts have been hampered by the sensitive redox chemistry of this system. 1.2. Structure of the Triarylmethyl Cation A more realistic picture of the binding site than 1 is provided by the crystal structure of the related cation tetrafluoroborate salt 62 which deviates from D3 symmetry in the lattice but can still be described as a propeller. The comparison of the X-ray structure (determined by Kahr and Blount at Princeton University) of the cation 62 and a calculated (MMP2) structure of the radical 1 is shown in Figure 10. As a consec groups are pockets wit? cavity. Figr state (fror 13. Stru Sim avoid son circumve ettmplexa dimethox boron rep tettowinp handled i Structure 96 As a consequence of the propeller conformations of l and 62 the methoxy groups are arranged in such a way that they form a pair of nucleophilic pockets with the oxygens' lone pairs projecting toward the center of a small cavity. Figure 10. Space-filling views of the binding site: MMP2 ground state conformation of radical 1 (left) and X-ray structure of 62 (from 62 ~BF4, right). 1.3. Structure of Triarylmethyl Borane Since small metal ions catalyze oxidation of 1 by 02, it is difficult to avoid some oxidation in all our preparations of complexes. Wishing to circumvent these difficulties while still gaining structural data on complexation, we have synthesized the diamagnetic compound tris(2,6- dimethoxyphenyl)methyl borane 63, a structural analogue to 1 in which boron replaces the central carbon atom. This compound is easily prepared following a modified Martin's procedure as shown below; it is easily handled in the absence of Brbnsted acids, it crystallizes easily, and its X-ray structure is shown in Figure 11, along with an MMP2 calculated structure of radical shown in I This neutrt in the solir barrier for structure r twists in tr crystal st dimethox: Fig ray 001'] 97 of radical 1 for comparison. A stereo view of the packing diagram is shown in Figure 12. \ . \0 / “L1, \\\\ ”"1 / 0\ 131:3-an 3b ———> O o O , \ This neutral isostructural analogue of 1 gives an approximate D3 structure in the solid state. From this structure and AMl calculations (rotational barrier for aryl rings; 4-6 kcal/mol),136 we conclude that the X—ray structure of the radical 1 is anomalous. The comparison for 'aryl ring twists in tripod binding sites are summarized in Table 1 and a summary of crystal structure determination and refinement data for tris(2,6- dimethoxyphenyl)methyl 1, tris(2,6—dimethoxyphenyl)borane 63, and tris(2,6-dimethoxyphenyl)methyl cation 62 are summarized in Table 2. Figure 11. Ball and Stick representations of the binding site: X- ray structure of borane 63 (left), and MMP2 ground state conformation of radical 1 (right). Figur It was $1 trimesitylb identical. 1 stmctural c cempound haem trierI’Ibor; difference; only a sir.- Significan Should 110‘ 98 Figure 12. Stereo view of packing diagram of borane 63. It was shown by Power that the solid state conformations of trimesitylborane and the trimesitylborane radical anion are virtually identical.137 The conclusion of this comparative study was that the structural consequences of the addition of one electron into the LUMO of a compound of the type Ar3X are trivial. While this conclusion is supported by a tendency toward D3 symmetry in many reported structures of triarylboranes, triarylmethyls, and triarylamines, there are gross differences in the crystal conformations of l and 62, which also differ by only a single electron. Similarly, the isoelectonic species 62 and 63 are significantly different. These results caution that simple generalization should not be drawn from a comparison of only two crystal stuructures. Tabl aTw soul solu Tal tris bor 99 Table 1. Aryl ring twists in tripod binding sites (in degrees)3 1 12.3 1b 47 1C 45 63 62.8 62 32.6 61.0 61.0 47 47 45 45 64.2 64.2 46.1 48.9 C2 Axis D3 D3 C2 Axis General site aTwists out of central atom plane; coplanar = 00° bDetermined in soultion by Martin using EPR measurements.C Determined in solution by Ishizu and Mukai using ENDOR measurements.138 Table 2. Crystal structure determination and refinement data for tris(2,6-dimethoxyphenyl)methyl 1, tris(2,6-dimethoxyphenyl) borane 63, and tris(2,6-dimethoxyphenyl)methyl cation 62. Crystal data 6 3 6 2 1 Space group C2/c Pl P2/n 2* 4 2 2 Temperature 295 1 10 295 a (A)** 11.076 7.214 10.405 b** 20.839 12.931 9.429 c** 9.944 13.633 11.767 or (°)** 90 83.13 90 B** 98.40 77.70 102.120 y** 90 80.56 90 u (cm-1)*** 0.80 (Mo Kor) 9.7 (Cu Koc) 0.83 (Mo Kor) *number of molecules in the unit cell, ** Cell dimensions, ***X-ray source. 1.4. Hex Tris dimethoxy various mt groups wl sterically ] position i substituen ether oxy ligand's bi Berane 6 63 with t methoxy Borane i ferrocen Voltamr tetrafluo Potentiai 00nditi0 100 1.4. Hexachloro Triphenylmethyl Radical and Borane Tris(2,6-dimethoxy-3,S-dichlorophenyl) borane 64 and tris(2,6- dimethoxy-3,5-dichlorophenyl)methyl radical 65 were synthesized with various modifications. The chlorine substituents are vicinal to the methoxy groups which form the ether tripod binding sites; they may therefore sterically perturb the methoxy groups, twisting them out of their preferred position in the aryl ring plane. These sites also position a vicinal substituent to withdraw or donate electron density by resonance with the ether oxygens, potentially modifying their Lewis basicity and hence the ligands binding abilities, without perturbing the radical center. Borane 64 show remarkable stability toward protic acid compare to borane 63 with better preorganization for ion binding because of the twisting of methoxy groups out of the binding cavities as shown in Figure 13. Borane 64 shows a reversible redox potential of -2.14V (referenced to ferrocene oxidation) for the process of B B" in cyclic voltammetry in methylene chloride with tetrabutylammonium tetrafluoroborante as a supporting electrolyte. Determination of the redox potential for borane 63 was unsuccessful under a variety of experimental conditions. Figr with ion- twis Fi so PC 101 Figure 13. MM2 calculated structure of borane 64 dimer with lithium cation showing improved preorganization for a ion-binding over that in borane 63 by twisting methoxy group twistings out of conjugation with the aryl rings. 0.000 I r trio‘s 2.000 1.000 _____ -__._-1.-_.__._I_.__-.L__-_L --_L__-._‘ ”.093 J l l l j .000 -1.600 ‘1300 -0.000 -0.400 -O 000 Figure 14. Cyclic voltammogram of 64 in methylene chloride solution of tetrabutylammonium tetrafluoroborate. The redox potential was referenced to the oxidation of ferrocene. 2. Subst It is scheme, tl the centra magnetic j therefore substitute The cyclir conveys 1 consists r unstirred this worl Saturated (Ag/AgC solvents Cttiomel 0t whic‘, OUT rest and met OXidatir Voltamr exeminr cumin 102 2. Substituent Effects. It is reasonable to suppose, based on our paramagnet coupling scheme, that variations in the odd electron's availability (spin density) at the central carbon and methoxy oxygen will affect the formation and magnetic properties of crystallized complexes of these radicals. We have therefore synthesized and examined the redox properties of various substituted analogues of the radical 1 using cyclic voltammetry (CV). The cyclic voltammogram is analogous to a conventional spectrum in that it conveys inforrnations as a function of an energy scan. Cyclic voltammetry consists of cycling the potential of an electrode, which is immersed in an unstirred solution, and measuring the resulting current. The potential of this working electrode is controlled by a reference electrode such as a saturated calomel electrode (SCE) or a silver/silver chloride electrode (Ag/AgCl). The accessible potential ranges for common electrochemical solvents have been studied and found to be -3.0 to 1.4 V vs SCE (Saturated Calomel Electrode ) for THF and -l.8 V to 1.9 V vs SCE for CHZCIZ, both of which are wide enough for our measurements.139 Our results show reversible oxidation and reduction waves for 1 in THF and methylene chloride as shown in Figure 15, indicating that all three oxidation states are stable, at least on the time scale of our cyclic voltammetry scans (50-200mV/sec). Various supporting electrolytes were examined, including tetraalkylammonium and alkali metal salts, in order to examine their effects on redox potential for ion-bindings. /t Fig kin rad res sec One stre‘ Cempone binding accessib mediates Valuable Choice ( electron focused 103 0.400 Figure 15. Cyclic Voltammogram of 1 in THF. Note that kinetically reversible electron transfer was found for both radical/cation and anion/radical couples (-0.49 V and -2.05V respectively vs. ferrocene oxidation), at scan rates of 200mV sec-1. One strength of our approach to magnetic materials is that the organic components can be modified in synthesis to adjust the redox and ion- binding properties of these multidentate radicals. As the range of accessible oxidation states is thought to be related to superexchange- mediated magnetic coupling in stacked materials, such control may prove valuable.1‘10,141 This tuning of the paramagnetic ligands, combined with choice of metal couplers, promises broad control over the magnetic and electronic characteristics of our extended complex materials. I have focused first on simple modifications to make derivatives of radical l, and then progressed to the more complicated systems. 2.1. Her Mar made mir approach targets, 1 procedure withdraw reagents. devoted : the desire Systemt Oxidatir detiyati Based cheract 104 2.1. Hexamethoxy Triphenylmethyl Radicals. Martin's synthesis of 1 is simple and economical. Although we have made minor modifications in building derivatives of radical 1, our approach is basically unchanged from the original work. In selecting new targets, I initially focused on substituents which can tolerate the synthetic procedures. However, many of the most potentially interesting electron withdrawing substituents cannot survive treatment with organolithium reagents. Such substituents are of particular interest and we have therefore devoted substantial effort to developing new syntheses in order to access the desired range of substituted radicals. \ O 1. PhLi or n-BuLi Cr2+ or —_> 2. (EtO)2CO . Zn/H30T or Vitamin C O / 6 2 1 Electrochemical studies on simple para-substituted triphenylmethyl systems have shown that the substituents can substantially alter their oxidation and reduction potentials.142,143 Functionalization of l to make derivatives similarly varies the redox properties of this family of radicals. Based on simple resonance pictures, the electron donor or acceptor character of 4-substituents in aromatic rings should strongly affect the energy 0 perturbin 105 energy of the singly occupied M0 (the SOMO) without significantly perturbing the oxygens of the methoxy groups. 1 :x =x'=x" = H 66 :x =X’=X" = Cl 67 : x =X'=X" = OCH3, OBu 68 : X =X'=X" = CH3 69 : x =X'=X" = Ph 70 : x = CH3 Table 3. Summary of electrochemical data for derivatives of radical 1 in THF a) and CH2C12 b). a) x X' x" R+ _ R' R' ..__—- R' H H H —0.49 V -2.05 V C1 C1 Cl -0.30 V -1.84 V OMe OMe OMe -0.89 V -2.29 V Ph Ph Ph 050 V -1.90 V THF Solvent; Reference: Ferrocene oxidation b) H H H -0.44 V -2.07 V C1 C1 Cl -0.34 V «1.84 V OMe OMe OMe -0.93 V -2.31 V CH3 CH3 CH3 -0.72 V ~2.19 V Ph Ph Ph 053 V -1.88 V CH3 H H -0.61 V -1.99 V CH2C12 Solvent; Reference: Ferrocene oxidation By using oxidation property. orbital en might fin helpful. and 70, s radicals while elr expected. are only condense The redr between 0f the 31 Where it (meta 01 P iS a r Peculia eorrelat Where ] unsubsr 144 106 By using electron withdrawing substituents in 1, we may produce oxidation-resistant derivatives of 1 which still retain 1's ion-binding property. . On the other hand, if triplet coupling is promoted by better orbital energy matching between radical and lithium cation orbitals, we might find donor or resonance delocalized substituents to be especially helpful. As a start on these studies, we have synthesized 66, 67, 68, 69, and 70, simple derivatives of 1. It is clear that electron donor substituted radicals 67 , 68, and 69 are more vulnerable to air oxidation than is 1, while electron acceptor substituted radical 66 is less so. As might be expected, yields vary from good to poor as some of the substituents studied are only marginally tolerant of the ortho lithiation conditions required to condense the intermediate triarylrnethyl carbinol. The redox potentials in Table 3 can be examined for a possible relation between half-wave potential (for the one-electron reduction) and the nature of the substituent. One such relation is expressed in the Hammett relation where k and k° represent rate or equilibrium constants for substituted (meta or para) and unsubstituted compounds, respectively. log k =p0' k0 p is a constant characteristic of the reaction series, and o is a constant peculiar to the substituents; cm is generally set equal to zero. The correlations of electrochemical data are made using the following equation where E1/2R and E1/2H are the half-wave potentials for the substituted and unsubstituted compounds. Note that in the equation, p is expressed in volts. 144 The intr electrode many or wave pc reductio cation/r: couples. Examin graphic (5 valur values becaus involvr correla correla cation induct contri' 001161: Worse times 107 AEI :EIR __E1H =p0' 2 2 2 The intrinsic difficulties of attempting E1/2 - o correlations when the electrode reactions are irreversible have been stated.”5 Unfortunately, many organic reductions and oxidations are irreversible and hence half- wave potentials cannot have thermodynamic significance. However, the reduction potential measured in this work are generally reversible for the cation/radical couples and reversible or nearly so for the radical/anion couples. Examination of a relation between half-wave potential and 0 may be done graphically merely by linear plotting of one against the other using known 6 values.”6 The correlations between half-wave potential and sigma values are not so good as we can see in the least-squares fit in Table 4 because of the complexities of reactions and conformational effects involved. It has been suggested that the cation stability should be correlated with 0+ values, whereas anion and radical stabilities are to be correlated with G or (5' values. 147 It appears as if substituents affect the cation with resonance and/or hyperconjugation terms in addition to inductive and polarizability terms, whereas only the polarizability contributes substantially to radical and anion stability. An attempt to correlate our experimental data with 6+ or (5‘ was unsuccessful and gave worse correlation parameters than with o (the values were summed three times for the trisubstituted radicals for the correlation). NV) , I93, EtV) I I -2 II -2-1 I . 108 E (V) 1 I ' Anion 0 Cation Figure 16. AE1/2(V) vs 6 in THF for HMTP methyl radicals a) in THF and b) in CH2C12. Table 4. Least-squares fit of Correlation for half-wave potential and sigma values for HMTP methyl radicals. AE1/2= pX-E'1/2 P E'1/2 #3 Rb (S (Cation)C 0.3969 -0.53 4 0.965 (S (Anion)C 0.3019 -2.00 4 0.852 G (Cation)d 0.3912 -054 6 0.905 (S (Anion)d 0.3177 -2.00 6 0.805 a Number of experimental points used. b Correlation coefficient. C THF solution. d CH2C12 solution. Bank pro radical sy simple su first of t] triphenyl balanced monomer are repor expect si function stabilizir maximal stabilizir unsubsti methyl 2 Bank s1 System Withdra effect; SUbStitu exactly Direct Variety commc 1imited 109 Bank proposed that substituent effects detected in CV of triaryhnethyl radical systems are not likely due to conformational changes but rather to simple substituent contributions. He recorded the largest change for the first of the three methyl substitutions for H in the para position of the triphenylmethyl system. Steric and conjugative effects are said to be balanced in each case to provide the optimum stabilities. For para monomethylated triarylrnethyl anion, the twist angles for the phenyl rings are reported to be 197°, 306°, and 448°, respectively.148 Thus, we can expect significant differences in the degree of conjugation of aryl rings as a function of substitution. For the electron-deficient cation the greater stabilizing effect can occur when the donor methyl group is in the maximally conjugated ring . Similarly for the electron rich anion better stabilizing effects are achieved when the maximally conjugated ring is unsubstituted. This analysis is based on the electron donor properties of methyl and the differential ring conjugation. Bank suggested that the preferred ring conjugation of the triarylrnethyl system should be determined by various substituents. In fact an electron- withdrawing group should reverse the pattern of sequential substituent effect; thus, for the anion, greater stabilization would result when a substituted ring is maximally conjugated. For the cation the prediction is exactly the opposite. Direct application of Martin's original synthesis of radical 1 calls for a variety of 4- and 5-substituted 1,3-dimethoxybenzenes. The range of commercially available dimethoxybenzenes is adequate if somewhat limited. In cases where the desired substituents could not tolerate Martin's procedure substituer assembler modified systems. functiona acid-stab breaking It has her amines c 7 2, do dimetho modifiee The desi With sod a Ihates COITeSp 111016 1' electroi 1 10 procedure, an alternative strategy would be desirable which would allow to substituents to be added after the triarylrnethyl framework was fully assembled. The resulting intermediate could then be converted to a new modified radical, opening the way to much broader class of substituted systems. With this goal in mind, we have explored the possibility of functionalization after assembly of the triarylrnethyl framework via some acid-stable derivatives in which the central carbon is sp3 hybridized, breaking up the extended rt-system. It has been found that ammonia, as well as various primary and secondary amines can add into the central carbon in cation 62. The ammonia adduct, 72 , does not have the proton acid sensitivities of tris(2,6- dimethoxyphenyl)methyl carbinol 71. This ammonium salt can be modified with a wide range of halides by simple electrophilic reactions. The desired radicals can be than generated readily by a simple diazotization with sodium nitrite in acidic water. In a less generally successful strategy, carbinol 71 can be reduced to the corresponding triarylrnethane 73 using sodium cyanoborohydride which is more robust toward both treatment with organometallic reagents and electrophilic aromatic substitutionsl‘t9 The corresponding hexachloro methane methylene yet found Although umnodifir evidently at deprot pathways 111 methane 76 was synthesized by a simple chlorination with S02C12 in methylene chloride. The products are well characterized, but we have not yet found a clean conversion back to the corresponding radical or cation. Although oxidation with p-chloranil is effective at converting the unmodified methane 73 to the cation, the hexachlorinated analogue 76 is evidently too deactivated to undergo the analogous oxidation. All attempts at deprotonation were also met with failure. A variety of less general pathways were also introduced for specific substitution patterns. 71 ‘ HBF4 \ 0 (3+ NaBH3CN Ceric Ammonium / 3 Nitrate 62 There ar can be u could ne availablr scheme. mainly carbona for furtl 112 There are several more reactions worth mentioning here. The methane 76 can be used for building several substituted triphenylmethyl radicals which could not be built by the standard procedure because of side reactions or an available precursors. One example is illustrated in the following reaction scheme. The methane 76 can undergo ortho-chlorine directed lithiation mainly by the induction effect followed by quenching with dimethyl carbonate to give various substituted triaryl methanes which can be used for further functionalization of aromatic rings in the system. W useful rr halogen unsatur: These r derivati radicals Electro have 1 precurr dirneth reactio 113 H-C While building these substituted radicals, two very interesting and useful reactions were somewhat serendipitously introduced: Electrophilic halogenations of resorcinol derivatives; and condensation of or, B- unsaturated ketone and dimethyl malonate followed by aromatization. These reactions, especially the electrophilic halogenation of resorcinol derivatives, provided several key transformations in building these radicals. Electrophilic halogen substitution and alkylation of resorcinol derivatives have been studied for decades. While pursuing the preparation of precursors for radicals, we were able to build a variety of halogenated 1,3- dirnethoxy benzenes which we found later in the literature.150 Some of the reactions are summarized in the following reaction schemes. / O\ /O O\ 2 Brg _____y Br Br X X X: H, Cl. 1, OMe, Ph / O O\ / O\ m C1 C1 X X X: H, C1, 1. OMe, Ph Most of 1 An ether amount ovemigh usually dimetho: dimethor The 00: followe phenyl. overall aldehy. aromat' 114 Most of the reactions are nearly quantitative with easily purified products. An ether solution of precursors was combined with a stoichiometric amount of halogenating agent at ~78 °C followed by slow warm up overnight and filtration of crystalline products. Washing with (1 usually gave clean crystals of the products except dimethoxybenzene which we could distill under vacuum dimethoxybenzene was synthesized with aqueous NBS. ’0 0~ ,o 0‘ Ct sozcr. Cr —*’ Cl 0 0‘ ’ U NBS ”U“ ——-> . Br ,0 o, ,o 0‘ U 4728‘ U C Cl Br iethyl ether 4-chloro . 4-Bromo The condensation of 0t,B-unsaturated ketones with dimethylmalonate followed by aromatization was a key reaction in the preparation of 4- phenyl substituted triphenyl methyl radical 70. With a good efficiency for overall reaction and wide commercial availability of a range of aryl aldehydes, these reactions have great potential for building various aromatic group substituted triphenyl methyl radicals. All of variah been i Figur deterr Tabb OO O D H A. NaOH MeoU OMe Naorat V All of the radicals generated have been studied by room temperature and variable temperature X-band EPR. Some ion-binding studies also have been carried out by EPR. Some of the typical EPR spectra are shown in Figures 17, 18, and 19 along with simulations. The experimentally determined coupling constants ai and the spin density pi are summarized in Table 5. 4—‘AJH-f‘ 11 f0 ra Kivelr tfiphm father subsd the m Garbo ava‘n; donat may : CatiOr 116 Table 5. A summary of 1H hyperfine constants ai (in Gauss). for coupling with meta and para proton in HMTP methyl radicals in THF. ameta (#)2 apara (#)2 Ipmetal1 tpparat1 1 1.06 (6) 2.26 (3) 0.047 0.100 14 1.07 (6) 2.35 (3) 65 2.3 (3) 0.102 663 .1.0 (6) 0.044 1 lpcxpl = la /22.5| 2numbers of protons in each positions. 3radicals 67 , 68, and 69 show similar results. EPR spectra simulated with the ESRa program written by A. K. Rappé and C. J. Casewit, Calleo Scientific Software, Colorado State University. 4Determined by Ishizu and Mukai using ENDOR.138 Kivelson reported that chloride substitution in the para position in triphenyl methyl increases the spin density at the ortho carbon slightly and rather more at the meta position.151 He suggested that methoxy substitution in the para position in triphenyl methyl radical does not affect the meta carbon spin density but increases the spin density on the ortho carbon considerably. We can follow the same arguments in estimating available spin densities on methoxy oxygen in our radicals with electron- donating para-substituents. The spin densities appear to be small but they may play a significant role in the magnetic interactions in our radical metal cation complexes. 117 Figure 17. EPR spectrum of l in THF with simulation (top). 118 —. Figure 18. EPR spectrum of 65 in THF with simulation (top). 119 ‘AA W v—v ' —A A.‘ W wfi ' Figure 19. EPR spectrum of 66 in THF with simulation (top). 2.2. Ti Sr undergo acidic c: the eye Figure cyclizer 120 2.2. The Cyclized Xanthenol Some of the cations which are intermediates in generating radicals undergo a cyclization to form a xanthenyl X (= OH, NH3+, N+H2R ) under acidic conditions by way of the mechanism outlined below. We believe that the cyclization occurs when two of the rings are forced to be coplanar. Figure 20 shows a stereo view of the X-ray crystal structure of the cyclized xanthenyl from carbinol 71. . \ o __\o o o o O C 70 GO :0 CD {<5 9* n 0 /\ ‘ H+ Figure 20. Stereo view of cyclized xanthenol from X-ray structure. 2.3. T 0 build lir tripod e ion-bin molecu packin; forms '. the twr a roug the so] 0 4/9 It 0 stab‘ com the dist 121 2.3. Tetramethoxy Triphenylmethyl Radicals Our purpose in using complexes of radical 1 with metal cations is to build linear chains of radicals by complexation of the metal ion with the tripod ether oxygens in the radical. Various ion-binding studies suggest that ion-binding can occur with as few as four methoxy groups in a given molecule. Furthermore, careful examination of the three-dimensional packing diagram of radical 1 shown below reveals that the radical itself forms linear chains without a metal cation. The four methoxy groups on the two more twisted rings are already placed in such a way that they form a roughly square planar ligand environment between adjacent radicals in the solid state. / , 2 .J. 79» ;“m/24664 ///// ‘/ It occurred to us that if a radical with only four methoxy groups would be stable enough to be handled in the usual way and could form a linear chain complex with metal cations between radicals, it should be possible to study the magnetic behavior of the resulting coordination chemistry. As I discussed earlier, the anomalous solid state structure for l is troublesome even tho anount thought defined altemati A seri methy modif usual derive dimet' Mukt Specu 65°, ; const Show const 122 even though various calculations show that it can be explained with a small amount of crystal lattice energy (3-6 kcal/mol by AMlcalculations). We thought that if the tetramethoxy radical could also give a structurally well defined complex, and magnetic interactions, this new complex might be an alternative building block for molecular magnetic materials. A series of para-substituted phenyl bis-(2,6-dimethoxyphenyl)phenyl methyl radicals was prepared as shown in next page by a simple modification of Martin's procedure. All of the radicals were studied as usual by EPR and CV as with the hexamethoxy triphenylmethyl derivatives; the results are summarized in Tables 6 and 7. The di-(2,6- dimethoxyphenyl)phenyl methyl radical 77 has been studied by Ishizu and Mukai using the Electron Nuclear Double Resonance (ENDOR) spectroscopy and was found to have C2 symmetry with twist angles of 65°, 65°, and 30° in toluene. They reported almost identical hyperfine coupling constants to our values for the radical 77 in toluene.152 This spectrum shows significant solvent dependence as seen from the hyperfine coupling constants in THF (Table 8). 123 77: X =X'= H 78: X = C1 79: X = OCH3 80: X = CH3 81: X = OH 82: x = COzH 83: X = N02 84: x = H, x: OCH3 Table 6. Summary of electrochemical data for derivatives of radical 19 in CH2C12. 4-X 3,5-X2' R+ = R R = R H H -0.32 V -1.90 V C1 H -0.30 V -1.84 V OCH3 H -0.53 V -2.04 V CH3 H -0.42 V -1.93 V OH H -0.58 V -1.82 V C0211 H -0.22 V -1.34 V N02 H -0.25 V -1.80 V H OCH3 -0.34 V -1.87 V CH2C12 Solvent; Reference: Ferrocene oxidation As with the l half-wave p01 (TMTP) met] correlations l complexities of hydrogen 81 and 82 sh: be explained It. Flgur in CH 124 As with the hexamethoxy triphenymethyl radicals, correlations between half-wave potential and literature sigma values of tetramethoxy triphenyl (T MTP) methyl radicals are poor, as shown in the least-squares fit of correlations in Tables 7; presumably, such difficulties reflect the complexities of reactions and conformational effects involved. The effects of hydrogen bonding in CV are not completely understood. The radicals 81 and 82 show the most dramatic departure from the series, which might be explained by the effects of hydrogen bonding and n-conjugation through it. ‘ o O -0.5'. o E(V) U -1.5" I I E] a I -25 v 1 f I ' I ' I fi I O4 02 00 02 04 06 08 Figure 21. AE1/2(V) vs 0' in THF for TMTP methyl radicals in CH2C12. Table potenti Bin 0 (Cat 0 (Am' a Nun c CH2 Some of the With mm the Spin der 125 Table 7. Least-squares fit of Correlation for half-wave potential and sigma values for derivatives of 77. AE1/2= pX- E'1/2 P E'1/2 #a Rb 0 (Cation)C 0.2957 -0.4029 8 0.790 G (Anion)C 0.2741 -1.8480 8 0.265 a Number of experimental points used. b Correlation coefficient. 0 CH2C12 solvent. Some of the typical EPR spectra are shown in Figure 22, 23, 24 and 25 with simulations. The experimentally determined coupling constants ai and the spin density pi are summarized in Table 8. Table 8. A summary of 1H hyperfine constants ai (in Gauss) for 126 coupling with meta and para proton in TMTP methyl radicals. aortho (#) ameta (111)2 apara (#)2 lporthI1 lpmeta|1 Ipparall 77 0.91 (4) 1.00 (2) 0.040 0.044 4 4.30 (2) 1.50 (2) 4.96 (1) 0.191 0.067 0.220 5 4.09 (2) 1.51 (2) 4,67 (1) 78 0.9 (4) 1.0 (2) 0.040 0.044 4 4.2 (2) 0.7 (2) 0.187 0.031 843 0.8 (4) 1.1 (2) 0.035 0.049 4 4.8 (2) 3.2 (1) 0.187 0.142 1 Ipexpl = |(a/22.5)| 2numbers of protons in each position. 3Radicals 79, 80, and 83 shows similar results. 4Asymmetric ring protons coupling constants in TMTP methyl radicals. 5Determined by Ishizu and Mukai using ENDOR in toluene.138 128 1 l O Figure 23. EPR spectrum of 77 in toluene with simulation (top). 129 Figure 24. EPR spectrum of 78 in THF with simulation (top). 131 2.4. Rotational Barriers Triaryl-X propellers, both planar and pyramidal, are well known to exhibit barriers to aryl ring rotation. Some of the triaryl systems constructed in this work and related studies show pairs of methoxy proton resonances at room temperature which are interpreted to indicate ring rotation that is slow on the NMR timescale. We thought that it would be useful to determine rotational barriers for relevant systems to have a better understanding of the structures of our systems in solutions. The rotation barriers of the aromatic rings in the 71 and 73 have been studied by Rieker and Kessler using 1H NMR and found to be less than 8.2 kcal/mol for carbinol 71 and 11.1 kcal/mol for methane 73.153 The silane 74 in which the central carbon in methane 73 is replaced by Si crystallizes on a three-fold axis so that all three rings are equivalent. There are two independent molecules in the unit cell and they have twist angles of 30 and 33° respectively. It is worth adding that the methane 73 crystallizes in the same space group as the silane and has similar ring twists (27°). Due to its single intramolecular hydrogen bond(one of the methoxy groups is twisted 75° out of the ring plane because of the hydrogen bond), carbinol 71 has a lower symmetry in the solid state (Pbca space group). Figure 26. X-ray structure of methane 73. 132 Variable temperature 1H NMR studies have been conducted over the temperature range of -90°C to +100°C using the Varian VXR 300 NMR instrument in toluene-d8 or CDC13 solvent to observe coalescence of two methoxy resonances in systems with doubled methoxy proton peaks at room temperature and therefore restriction of free rotation of rings. The results are consistent with a site exchange process resulting from the rapid flipping of the aryl rings on the NMR time scale. The value of the rotational barrier has been determined by applying the Gutowsky-Holm approximation to the observed methoxy group site exchange at the coalescence temperature (TC in K).154 For an exchange process between two nuclei A and B with a mutual coupling J AB with peak separation of Av in Hz, the rate constant kC at the coalescence temperature To is given by: kc = 2.227sz + 6J2” From the Eyring equation, one can determine the free energy of activation by using the following equations. To calculate the activation barrier, we need a pair of kC and Tc. k3 is the Boltzmann constant, K is a transmission coefficient usually assumed to be 1, and h is Planck's constant. @e—Ac‘mr k=K AGf = 4.58TC (10.32 + log—£90m] / mol C AGC¢ =19.14TC(10.32 + log £9” / mol C 133 The activation barriers, AGi , for aryl ring rotation of tris(2,6- dimethoxyphenyl)methy1 ammonium tetrafluoroborante 72, tris(2,6- dimethoxy-3,5- dichlorophenyl) methylammonium tetrafluoroborante 75, and tris(2,6-dimethoxy-3,5- dichlorophenyl) methane 76 were 12.67i0.3 kcal/mol-l, 15 .7i0.8 kcal/mol, and 17.3i0.8 kcal/mol, respectively. Table 9. Variable Temperature NMR results for tris(2,6- dimethoxyphenyl)methy1 ammonium tetrafluoroborate 72, tris(2,6-dimethoxy phenyl)methane 73, tris(2,6-dimethoxy- 3,5-dichloro phenyl) methylammonium tetrafluoroborate 75, and tris(2,6-dimethoxy-3,5-dichlorophenyl) methane 76. 72 73a 75 76 ASH (CH3O, ppm) 0.63 0.254 0.331 (at T/K) 233 --- 263 253 AGi (kcal mol'l) 12.67 11.1 15.68 17.27 Tc (K) 273 216 323 358 kC (8'1) 419 —-- 169 220 AV (Hz) 189 33 76.2 99.3 3Determined by Rieker and Kessler in CDC13. 134 .____.. .__, ,,_ -\ A .__- h”--. -_- _._.—-. -_—-__ 17'"[W'IllllI—rlllllll'IlYlllllll'lTllITlllllllTlllllllllllllll’llll‘llT—q 4.4 4.: 4.- 4.1 4.0 71.9 3.8 3.7 3.b 3.5 3.4 3.3 .3,; 00m Figure 27. 300 MHz 1H NMR signals of the methoxy protons in hexachloro tris(2,6-dimethoxyphenyl)methyl ammonium BF4 75 , recorded at different temperature. For slow rotation two peaks are obtained, whereas for fast exchange there is only one. In the intermediate range the signals are broadened. The coalescence temperature Tc is 50°C. (-10, 20, 30, 50 °C from the bottom) 135 _ ._»\ ._ , .4»w~»—'§_—-W‘r (x A ’-._’ 'V 7 ‘ d“ WM .IVWVIY11117111[TllllllTTfTllTlll'llllllllT‘IIlllTITllllllllllIllT—Tlllll111111111111111171'111 .- 4.0 3.8 30 34 3.2 30 2.8 2.6 2.4 Figure 28. 300 MHz 1H NMR signals of the methoxy protons in hexachloro tris(2,6-dimethoxyphenyl)methane 76, recorded at different temperatures. For slow rotation two peaks are obtained, whereas for fast exchange there is only one. In the intermediate range the signals are broadened. The coalescence temperature Tc is 100°C for the methane. (20, 40, 60, 80, 100 °C from the bottom) 136 Unlike 72, 7 5, and 76, HMTP carbinol 71, and HMTP silane 74 do not show splitting of methoxy 1H resonances at 300MHz even at -90 °C. These results are consistent with a faster flipping of the aryl rings than NMR timescale even in carbinol 71 in which we believe that there is a hydrogen- bond between one of the methoxy oxygens and the hydroxy proton at the temperature. We may understand why the carbinol 71 has a lower barrier for rotation than the methane 73 using "the two-ring-flip mechanism" of Mislow. The formation of the hydrogen-bond would fix the twist angle of ring in a "vertical" orientation, making it easier to rotate the other two rings appropriately in carbinol 71. By assuming the methane 73 as a worst case of all the triaryl methyl systems studied as far as rotation barrier of rings (except the case of 72, in which the methoxys are twisted), it would be safe to say that the activation barriers for the rotation in these systems are smaller than 11.1 kcal/mol in solution of CDC13 and toluene-d8. The rotation of rings in 71 and 72 would result a cleavage of the hydrogen- bonding interactions. We expected that the rotation barrier for methoxy groups in these systems will reflects the strength of these interactions. The carbinol 71 does not show any splitting of the methoxy proton signals at -90°C in toluene-d8. The difference between the rotation barrier of 72 and 73 is only 1.6 kcal/mol which might reflect the strength of these interactions. It is difficult to understand why they show such a small difference when 72 has three hydrogen—bonding interactions in the solid state compared to 73 which does not have any. 137 3. Ion-Bindings Operating on the supposition that the relaxation of 7Li nuclei (S=3/2) would be considerably shorter in the complex l-LiBF4 than in uncomplexed mixture, Jackson and Kahr compared the line widths of the solid state 7Li-NMR spectra for the precipitated solids with pure LiBF4 salts. The 7Li line widths in the reported complexes were 16 kHz at half height as opposed to 500 Hz in the pure salts. Jackson and Kahr interpreted these changes in terms of a specific binding interaction between radical l and the 7Li ions in the precipitated solids; however these results do not exclude other nonspecific broadening mechanisms. 3.1. Borane Small metal cation binding studies of borane 63 by UV—Vis spectrophotometry suggest ion-binding, consistent with the notion of stack formation. Similar UV—Vis studies of 1 are hindered by the Li+ catalyzed oxidation reaction which yields the intensely colored cation 62. \ o O/ LiX UK a a B 0 Hz0 < O I 138 UV-Visible spectrophotometry provides a useful probe of the binding of small metal cations by radical 1 and borane 63. The long- wavelength maxima in each of these species in ether, THF, CH2C12, and CH3CN are markedly diminished in intensity by addition of solutions of lithium salts (LiBF4, LiClO4, LiSCN). Washing the complex solutions with water recovers the original species' spectra. Analogous spectral changes are not seen in solutions of anisole or m-methoxyanisole upon treatment with lithium salts; the only observed changes correspond to simple dilution of the sample solutions. We interpret these observations in terms of the binding of small metal cations to the triaryl propellers l and 63, which alters their absorption spectra. These ligands are apparently not such efficient ion binders that they can compete with bulk H20; hence the observed reversibility upon washing as shown in Figure 29. Ion binding studies by NMR on the borane using various Li and Na salts in various solvents do not show significant changes except in the case of Lil-Borane which is still small (0.02 ppm changes under very anhydrous conditions) presumably because of the very low binding constant involved. ~—-r1 139 1. 500 .37 \ /\ l ,«" \.. 1‘ ,-' \ ABS //"*\\/" ‘ \J’ \ \3 3.000 A—C—uu __________ a ABS 3. 000 fl , j _% \ c) ABS 0.000 , , , 4 l rm 240 mo 320 360 400 Figure 29. UV-Vis spectra illustrating reversible Li+ binding in CHZClz. (a) Borane 63 alone; (b) Mixture after addition of LiBF4 solution; (c) Organic layer after H20 wash. . 140 The FAB Mass spectrum give one more piece of evidence for a notion of binding of radical 1 with metal cation. It shows RMR peaks in the case of NaBPh4 complex of radicals 1 and 68 which means that the complex formed by simple mixing of salts with radical is strong enough to remain intact through the FAB ionization process even if it only shows very little changes in other spectroscopy. J502180005 Scan 2 HT=0:38 100%:410000 um 19 Feb 93 14:54 Compacted SLRP +EI SA70(NBA)1 + NaBPh4 100— 446.0 151.0 R'——> ‘— R'Na‘ 523.0 04 -AL _. i A 5"0 ultrrftllilluriirrrrr 150 200 250 300 350 '400 450 500 550 600 650 1007 .1 I R'Na‘R’ ”0.2 m2 ' 0311.2 1045.2 1130.5 1193.5 o—J1110721llIllllllll’llllrlll‘llljj 700 800 900 1000 1100 1200 Figure 30. Mass Spectrum of dimer 2 of l-Na+-1 (869) in m-nitrobenzylamine matrix. 141 3.2. Double faced Paramagnetic Ionophore Substitution of Z for the methyl groups shown below in 1 allows for a whole range of new radicals to be synthesized. By controlling the size, electron withdrawing/releasing properties, and ligand capabilities of Z, the ion binding properties of 1 can be varied. Synthetically, the Z-sites are certamly the easiest to substitute; resorcinol (3-hydroxyphenol) and 3- methoxyphenol are commercially available starting materials, and Williamson ether synthesis provides a straightforward means for attaching many possible 23. The theme of two radicals sandwiching a metal ion can be inverted in a complex with two metals fixed around a single radical center. To evaluate the validity of our electron-coupling strategy, it would be useful to know the dimensions and the spin density on the bound metal cation. From such data it should be possible to estimate limiting magnitudes for the exchange coupling which could occur through the metal ion couplers in the hypothetical structure. 142 86-2LiBF4 85-1.5LiBF4 We have recently reported155 the radical 86, a double "octopus" analogue of 1.156 This work addresses the spin density questions and simultaneously provides structural insight into the novel two-faced binding properties of the hexamethoxy-triaryl complexants. In parallel with the studies of 86, radical 85 has also been built, with the aim of making a radical with one face closed to chain formation. Radicals 85 and 86 were synthesized by straightforward modification of Martin's original procedures. Unlike 1, compounds 85,86, and their related triarylrnethanol precursors are oils. Purification in this system is most conveniently achieved by crystallizing tetrafluoroborate salts of the corresponding triarylrnethyl cations; these are subsequently reduced to make the neutral radicals. CW-EPR spectra collected for 86 at -70 °C showed a 13- line pattern with the splitting between consecutive lines being 1.0 gauss. Figure 29 shows the experimental and simulated CW-EPR spectra of 86. 143 UJU Figure 31. EPR spectrum of 86 with simulation (top). 144 Our purpose in synthesizing 86 was to build an analogue of 1 which could not form extended chains by pairing around metal cations as in dimer 2. Radical 86 offers a bound Li+ ion a full coordination sphere of six ether oxygens, effectively "capping" the radical on both faces. Thus, the radical center in 86-2LiBF4 is encapsulated and should behave as a completely isolated paramagnet. Like 1, radicals 85 and 86 show some tendency toward oxidation on treatment with small metal cations in air. This behavior is evidenced by the appearance of the blue color of the corresponding triarylrnethyl cation in each case. The electron spin echo envelop modulation (ESEEM) technique of pulsed EPR spectroscopy has been used to measure weak electron-nuclear hyperfine coupling between the radical centers and coordinated metal cations in 86 in collaboration with Professor McCracken and Mr. Hong-In Lee. The ESEEM method provides structural details regarding radical- metal ion distance, the number of metal ions held near a given radical center, and an estimate of the unpaired electron spin density transferred to the metal. 145 ‘0 A f A U I A U L 1 J L l 1 J 0246,8101214161820 froquoncy 0.1:) Figure 32. ESEEM spectrum of 86-2LiBF4 in 1:1 THFzToluene at 4 K. The spectra are fourier transforms of time domain data collected under the following conditions: microwave frequency * 9.346 GHz; magnetic field strength 3340 G; microwave pulse power 50W; pulse widths 16ns FWHM; two pulse excitation. L_“ 146 Addition of excess (10 equivalents) of LiBF4 and NaBPh4 to 10'6 M THF solutions of 86 results in complex 86-2LiBF4 and 86-2NaBPh4. Because no new hyperfine couplings could be resolved in the CW-EPR spectrum after this treatment, the ESEEM technique was used to check for possible weak 7Li hyperfine coupling to the paramagnetic center.157 ESEEM data were collected on a home-built spectrometer described elsewhere.158 A two-pulse (900-1-1800) ESEEM spectrum collected at a microwave frequency of 9.35 GHz and a magnetic field strength of 3350 gauss is shown in Figure 32. The prominent doublet centered at 5.6 MHz shows a splitting of 1.0 MHz. When ESEEM data were collected at 8.52 GHz with a field strength of 3052 gauss, the center of the doublet shifted to 5.0 MHz, consistent with its assignment to 7Li. Also present in the spectrum of Figure 32 are 7Li sum and difference combination frequencies centered at 11.2 and 1.2 MHz, respectively, and a peak near 14.3 MHz due to weakly coupled protons. An analysis of modulation frequencies and depths using the formalism of Shuben and Diknov showed that this coupling arose from two 7Li+ ions coupled to the radical center at an effective dipole-dipole distance. of 3.5 A and having an isotropic hyperfine coupling of 0.4 gauss.159 Because the line width of the sum combination peak is independent of hyperfine anisotropy, the amplitudes and damping factors for the fundamental peaks (5.0 and 6.2 MHz) relative to that of the sum combination line can be used to determine the dipole-dipole distance. 147 % k A L L A‘ .A A L é 4 6 a :0 frequency (MHz) Figure 33. ESEEM spectrum of 86-2NaBPh4 in 1:1 THFzToluene at 4 K. The spectra are fourier transforms of time domain data collected under the following conditions: microwave frequency 8.90 GHz; magnetic field strength 3195 G; microwave pulse power 25W; pulse widths 16 ns FWHM; and a 1: value 147ns for a stimulated echo excitation sequence. 148 The absolute intensities of the lines can be used to determine the number of nuclei that give rise to the coupling. No coupling is seen to 10B, 11B, or 19F nuclei of the BF4" counterions; thus the splittings do not simply arise via nonspecific coupling to magnetic nuclei near to the radical. Analogous studies of 86-2NaBPh4 and 86-2Nal complexes gave similar findings; a stimulated echo spectrum of 86-2NaBPh4 is shown in Figure 33. Two peaks centered at 3.6 MHz, the 23Na larmor frequency, and split by approximately 2.5 MHz are observed. The lack of a pronounced sum combination peak in the two pulse ESEEM data precludes independent determinations of dipole-dipole distance and the number of coupled Na+ ions. Computer simulations of these ESEEM data are consistent with an isotropic hyperfine coupling of 2.4 MHz, and an effective dipolar distance of 3.8 A. Values for reff are calculated assuming that two Na+ cations are coupled. If only a single Na+ nucleus is coupled the calculated dipole- dipole distance decreases to 3.3 A, a distance that seems unlikely based on CPK models and MNDO studies. Assuming that the 23Na hyperfine coupling constants can be compared with their atomic values, unpaired spin densities of approximately 0.3% can be estimated for both Li and Na.160The Li+-radical center distances from ESEEM analysis using the point dipole-dipole approximation are in good agreement with those obtained from MNDO calculations. These results are summarized in Figures 34, 35, and Table 10. A larger splitting of 6.0 MHz is also observed in the Na salt-treated samples of 86. This result strongly suggests a complex with only one Na+ 149 bound, inducing pyramidalization at the radical center and hence decreased distance and increased overlap between radical center and N a+. Figure 34. Summary of ESEEM spectroscopic and MNDO calculated results. CPK models yield similar geometrical characteristics. / /% 6% )- i' ‘I rig $7.}; 051 y/j/x. ./,,. . Figure 35. Stereo view of the MNDO calculated structure; the BF4‘ counterions were left out of the calculations. Table 10. ESEEM Determined Complexation Data for 86 with Various Salts. Salt AisoaVle) rem/1) IMNDO #bound M+ Spin Density on W LiBF4 0.9 3.5i0.2 3.45 2 0.3%! LiI 1.0 3.4i0.2 3.45 2 0.3%a NaBPh4 2.4 38:20.20 b 2 0.3% NaI 2.4 3.8i0.2C b 2 0.3% a. MNDO calculations used starting geometries generated from a simpler MNDO calculation on 1-2LiF by "growing" the ether atoms in the expected lowest energy conformation, as evaluated using the Chem 3D molecular mechanics package. Optimization was carried out in D3 symmetry, as suggested by calculation on 1, the above complex, and related systems. MNDO calculates a spin density of 0.22% on the bound 7Li ions, in reasonable agreement with experiment. Unfortunately, without suitable reference systems, we cannot judge the significance of this value. b. MNDO parameters are not available for Na so these structures were not calculated. Attempts made using "sparkles" which represent Na+ as a positively charged hard sphere gave unreasonable structures. 151 Electrochemical characterizations of 85 and 86 were accomplished via cyclic voltammetry (CV)161. The salts 85+BF4— and 86+BF4-(1 x 10-4 M) were studied in THF and methylene chloride solution using tetrabutylammonium tetrafluoroborate as a supporting electrolyte. Reversible reduction of each cation to the corresponding radical was readily measured; a second reduction wave (radical to anion) could also be 1 observed when THF solutions of radicals were used instead of the cation salts. Table 11 compares the observed reduction potentials for l, 85, and 86. Table 11. Summary of electrochemical data for radicals 1, 85, and 86 with data for radical 1 for comparison. in THF a) and CH2C12 b). a) R+ :=——‘ R. R. 2 R- l -0.49 V -2.05 V 85 -0.47 V -2.10 V 86 -0.48 V -2.04 V THF Solvent; Reference: Ferrocene oxidation b) 1 -0.44 V -2.07 V 85 -0.50 V -1.98 V 86 -0.48 V -l.97 V CH2C12 Solvent; Reference: Ferrocene oxidation 152 As expected, the redox potentials for 1, 85, and 86 are nearly identical, which shows that the added ether extensions in 86 have essentially no effect on the redox properties of the Ar3C nucleus. The separations between cathodic and anodic peaks were large compared to the theoretical limit for Nemstian behavior (0.06 V for peak to peak distance); the wave for ferrocene/ferrocene+, a well-behaved reversible redox couple, showed similar separations. Thus, we believe electron transfers from an electrode to Ar3C substrates are rapid and reversible. In principle, electrochemistry should be a useful probe of ion binding in triaryl-X systems such as 1 and 86. Placement of a metal cation close to the center of radical 86 would be expected to broaden and shift both reduction waves to more positive voltages. These expectations are qualitatively borne out, but the voltammograms obtained are so poor and irreversible that we cannot identify well-defined reduction potentials for the metal cation-containing complexes. In support of the idea that specific binding is occurring, we note that no changes are observed in the cyclic voltammograms when salts of larger cations such as Na+ or K+ are substituted for the Li+. 153 3.3. Tetramethoxy Triphenyl amines For the purpose of ion-binding studies of derivatives of radical 77, structurally analogous amines in which the central carbon is replaced by nitrogen have been synthesized. In NMR ion-binding studies, it has been found that the use of metal salts with non-reducing counter anions such as BF4- and C104; generate significant amounts of radical cations of triaryl amine which can complicate data analysis by broadening NMR lines. However, Scott Stoudt has demonstrated that use of I' salts effectively protects against the oxidation process of the triarylamines; I- is easily oxidized and evidently is sacrificed to make 12 instead of allowing the amine to be oxidized to make the persistent amine radical cations. Thus, in the case of LiI we were able to study ion-binding in CDC13. Furthermore, iodide also appears to be a strong enough ligand in organic solvents that it "caps" the tripod-bound lithium cation, filling out its tetrahedral coordination geometry instead of giving way to allow a second tripod ether to make an octahedral cavity. The amines 87 and 88 shown below are isostructural to radicals 78 and 84. They were synthesized by a modification of Fréchet's method for building triarylamines as shown in next page.162 These amines were used in ion-binding studies with metal cations to demonstrate the abilities of four methoxy groups as coordination ligands. 154 \o X' 0— ‘ 1 N X \.. N X' 0 0— ’0 872 X = C1, X'= H 89 88: X = H, X'= OCH3 X I X' \0 X' X' D ’ O\ c K (:0 ‘ — u. o L + 2 I, N((CH2)20(CH2)20Me)3’ X D N 0 NHz X' IO 2 They show interesting binding properties which can be detected by NMR. Similar to the ion-binding experiments with borane 63 and radical 1, these triaryl amines can be used for more quantitative studies. Addition of salts to CDC13 solutions of amines in NMR tube results in significant chemical shift changes (0.2 ppm); D20 washing regenerates the original spectrum of triaryl amines. Figures 36 and 37 show 1H and 13C NMR spectra for such experiments using amine 87. Other salts including Na and Cd did not induce changes in NMR spectra. Figure 38 shows an ion-binding study of amine 88 by 1H NMR which has internal reference built into it. Our expectation for ion—binding of tetramethoxy derivatives and dimethoxy benzene suggest that the two methoxy groups in the 3,5 positions should not participate in ion-binding. To our surprise, even those two methoxy groups show chemical shift changes after adding Lil to the CDCl3 solution of amine 88. These ;— 155 findings may reveal conformational changes by ion-binding and suggest that the tetramethoxy amines in solution may not be preorganized for ion- binding. The possibility of participation of the central nitrogen atom's lone pair electrons is still an unresolved question. The radical cations of these triarylamines are isoelectronic to radicals 79 and 84 and show interesting redox properties. Those radical cations might offer alternative building blocks for magnetic chains with the required stability and structures. Their redox properties were therefore examined by cyclic voltammetry. Electrochemical data are summarized in Table 12; Attempts to conduct ion-binding studies by CV were unsuccessful due to the fact that the oxidations of these amines were not reversible. Table 12. Summary of electrochemical data for oxidation of ——__¥ +. . . . N ‘———— N in amrne 87, 88, and 89 1n CH2C12. 8 7 +0.50 V 8 8 +0.44 V 8 9 +0.57 V CH2C12 Solvent; Reference: Ferrocene oxidation 156 3CL40ME AMINE +LII 1 020 —7.2400 16.5628 15.5351 JUL _ -___L -- -- L _ 11. 10 a - -1, . in 15' t , TTTTTTT TI T I I T I T I I Y I T I I. I Y I V . I I T 6.0 5.5 5.0 4.5 4.0 3.5 PPM ' 1 3.0 Figure 36. 1H NMR spectra for the complexation of 87 with Lil in CDC13 showing chemical shifts of methoxy and corresponding aromatic protons followed by D20 wash. 157 W rmr" " '1" """ 1"7'I'T'Yfirer'V‘l'V'l VIII I'II Irrv vva IIIrrIIII 1111 1'11 rITI n” 1+1...fir 160 151) 140 130 1211 110 l 101) I 911 811 I 70 I 56 9911 Figure 37. 13C NMR spectra of 87 with Lil in CDC13 showing chemical shifts of methoxy and corresponding aromatic carbons. 158 <— 3,5-OMe 2,6—OMe —>- L A LL (1 . .- - - .. _ - -_.. __.-__ -.._... --... L..._______ L__ _ LA ii _ 1L___ __ L_ [7W v , . . r f . - ‘7 H I 1 l "'1»"'i""1"7 I‘ll‘fl—YTIY1‘. ;r: 11111 .fit.III‘.I1111.‘V.1'Itr!'v.‘.. 10 9 8 7 6I 5 4' 31 I 2' I 1l PPM Figure 38. 1H NMR spectra for the complexation of 88 with Lil in CDCl3 showing chemical shifts of methoxy and corresponding aromatic protons. ———— 7 F W| .. . . I * ‘-.vr|r*rw.;rfrrl7rIrrrT1I IITfifi.Y—fl..firIII—r—frlf.IIj—frrfi—i—rf7—flfi7—'fi‘.fi'—;—r I 800 100 160 140 120 100 80 60 4o 20 PPM 0 Figure 39. 13C NMR spectra of 88 with Lil in CDC13 showing chemical shifts of methoxy and corresponding aromatic carbons. 160 3.4. Ammonium Complexes by Hydrogen bonding As a way to construct desirably substituted radicals like perchlororinated triarylrnethyl radicals, we were working on the route for making the radical through tris(2,6-dimethoxyphenyl)methyl ammonium tetrafluoroborate 72 which is not as sensitive to protonic acids as the corresponding carbinol and can be converted into the radical by a simple diazotization reaction. The X-ray structure of the given ammonia adduct 72 shown in Figure 40 shows three hydrogen-bonding interactions, which hint at the hydrogen bonding ability of the ether tripods in radical 1. An attempt has been made to determine the ability of radical l to hydrogen bond with ammonium salts by using the borane 63 as a model compound in CD3CN as a solvent. To our suprise, after adding the salts into a solution of the borane 63 in a NMR tube, the characteristic proton peaks (equal-intensity triplet from coupling to N) of NH4I appeared which corresponded to eight protons per molecule of borane from the integration of the peaks. Intrigued by this fact, we tried a similar experiment on hexachloroborane 64, resulting in similar behavior except that in this case we can assign only four ammonium protons per molecule of the borane 64. The X-ray structure of the hexachlorinated ammonium salt 75 has recently been solved; the aryl ring 161 twists have turned out to be 37, 51, and 51°. As we expected, the methoxy groups in 75 are twisted down out of the aryl ring planes, improving preorganization for binding; this twisting (the average of 67°) can be seen in the following stereo view of the X-ray crystal structure of 75. a) b) Figure 40. Stereo views of crystal structure of a) tris(2,6- dimethoxyphenyl)methyl ammonium tetrafluoroborate 72 b)tris(2,6-dimethoxy-3,5-dichlorophenyl)methyl ammonium tetrafluoroborate 7 5. Note that at the time of writing, the ammonium H atom positions had not been finally refined. 162 Table 13. Crystal structure determination and refinement data for tris(2,6-dimethoxyphenyl)methyl ammonium tetrafluoroborate 72 and tris(2,6-dimethoxy-3,5—dichlorophenyl)methyl ammonium tetrafluoroborate 75. Crystal data 7 2 7 5 Space group P2/n P21 /n Z 4 4 Temperature 296 K 298K a (A) 15.292 12.03 b 10.796 24.07 c 15.731 12.1 1 [3 109.99 94.38 11 (cm-1) 9.7 (Cu K01) 0.8 (Mo K01) 163 14321 .:;;.223g - 1.9299 .-351 ..217 ‘, b54911; L_ r” \L 1 o I 'l;"% 1 1 T I [W 117 I'fiIT ‘rrVV'fiIFV 1 ‘0 9 H I s 5 4 3 2 .J 115 26.0 ”'9 P"; Figure 41. 1H NMR spectrum of complex of 63°NH4I in CD3CN showing characteristic triplet peaks after the addition of the salt which was not present before. 164 33801— (3. CI) I 9299 0 It. rx 1 —- — —-~—-—————-—-— - -~»—_--l-h—n-—J - IIIIIIIjIIWW'IIII I II IIIIIIIIII 8 7 2 1"filr"' m I I WTYUTI'fi'j—Iljlmn‘rtiUT'IWT[YrT_Y_ 6 5 4 _3 19PM '—1 5—3——J ' 14.1 I .4 1%.! —1 0"I‘ITTIIIIIrrTrI Figure 42. 1H NMR spectrum of complex of 64-N H41 in CD3CN showing characteristic triplet peaks after the addition of the salt which was not present before. 165 3.5. Metal Ions A variety of new salts have been investigated in reaction with 1 and the related radicals described above. We have examined a variety of salts, varying both metal cations and their anion partners. Solid phase SQUID and EPR studies at low temperatures were performed on all complexes, giving detailed information about the magnetic properties of these materials. As with X-ray diffraction, single crystal studies would be particularly valuable, as they would provide data on the magnetic anisotropies in such crystals. Howerver, among the magnetically interesting materials only polycrystalline solids were obtained. Dirners and oligomers of types 2 and 3 can obviously be envisioned with metal ions other than lithium. In addition to the few salts that we have already investigated, we have examined various transition metal ions, including Cr3+, Co3+, Cd2+, and Rh3+, all ions with strong affinities for ethers and with coordination spheres of appropriate size to bind to the ether tripods in 1. In a qualitative survey of roughly 20 alkali metal, alkaline earth, and transition metal salts, we examined the color changes due to radical oxidation in the presence of metal salts and Oz. A clear trend emerged, with metal ions of small radius (S 1.1 A)163 promoting the rapid color change and larger ones failing. These observations are consistent with the observed size of the binding pockets in 1 and 63 and we expect such survey data to be useful in helping to select new salts for further study. 166 The counter anions of the salts used in our complexation studies may be critical to the successful formation and isolation of crystals for X-ray studies. Crystallization of the extended (radical-metal)n stacks that we envision may be constrained by the requirements of fitting together such infinite charged structures in an efficiently packed lattice. Furthermore, the metal ions' charges are spaced at fixed intervals along such stacks or chains, as determined by the cation's radius and the aryl ring twist angles in the radicals. To maintain charge neutrality, the counterions must find appropriate sites with the right spacing to allow the chains to pack together in one of the relatively few simple patterns available to them. Counterion size, shape, and charge may therefore play a key role in determining which systems of salts and radicals will lead to well-formed crystals. 167 CHAPTER II. Magnetism of Triarylmethyl Radical Complex 1. Calculations on the Metal-Mediated Pairwise Interaction of Methyl Radicals Professor Jackson has examined and compared the total and one- electron frontier orbital energies of calculated singlet and triplet states for the simplest model system, H3C-Li+-CH3. The linear interaction of two carbon 2p-orbitals on radical centers, communicating via the 2s and 2p orbital set of a lithium cation, can be described by a simple qualitative interaction diagram, shown in Figure 43. . A10 H3C° LI+ °CH3 Figure 43. Qualitative MO diagram schematically indicating the coincidental near-degeneracies from mixing a pair of sp3 hybrid orbitals with a Li2pz orbital and a LizS orbital. 168 Preliminary work on the metal-mediated radical pair H3C~Li+oCH3 (3-21G and 6-31G*), suggests a triplet ground state for this model system. Geometry differences between singlet and triplet optimized structures are small, and vibrational analysis on the triplet shows that the linear form is a local minimum. The fact that the calculated total energies for the two spin states are close is surprising, and we need to pursue more complete wave functions which will determine the ordering and separation of the two Spin states as a function of C-Li distance. In conjunction with our molecular mechanics and semiempirical molecular orbital modeling of the actual radicals under study, we expect these higher-level calculations ultimately to lead to a set of generalizations which will help guide the design of new paramagnetic complexants and extended structures. Techniques such as MNDO, AMl, which are widely available and easy to use, provide a reasonable middle-ground for attempts at studying large systems such as radical 1 and its complexes. Reliable absolute predictions of preferred electronic couplings might not be possible using these methods, but they should help to predict trends and guide our choices of modified systems. Professor Jackson has fully Optimized the structure of radical 1 by MNDO and AMl within D3 symmetry, and its structure is nearly identical with that obtained by the molecular mechanics methods using PCModel. The program PCMode1164 offers molecular mechanics calculations based on MM2 force field. Calculations using the program have been shown to work well in predicting geometries for the triaryl-X monomers whose structures we have studied by X-ray. As a rapid source of reasonable estimates for geometries and strain energies, this program is a valuable tool. 169 2. Antiferromagnetic Interaction and Peroxy dimer of The Radical 1 Magnetic susceptibility measurements on crystals of radical 1 show very weak antiferromagnetism according to Curie-Weiss behavior. It is not clear what is the mechanism responsible for the coupling, but careful examinations of the three-dimensional packing of the radical show a closest distance of 4.20 A between para carbons of the flattened conjugated aryl rings of the radical. Curie-Weiss Fitting of HMTP Methyl Radical 200 y = 8.807] + 3.1449x R"2 = 1.000 0 =-2.8K J=2k0 = -7.73e-23 J =-3.88 cm-l “X 1001 - ' I ' I ' 0 10 20 30 40 50 TEMP (K) Figure 44. Curie-Weiss fitting of HMTP methyl 1 with Curie- Weiss constant of -2.8K (J = -3.88 cm'l) suggesting weak anti ferromagnetic interactions between radicals. Radical 1 reacts slowly in the air to form peroxide 90 (shown on next page) as determined by 1H NMR, FAB mass spectrometry, and elemental analysis. The formation of peroxide can be explained by the 170 reaction of a spin-rich para-carbon in the radical 77 with oxygen. According to McConnell's approximation there is a significant amount of unpaired electron spin density on that para-carbon which might be available for the reaction. Since it is known that oxyradicals add to p- positions of triarylrnethyl radicals, this reaction is the point of initial oxygen addition. The tetramethoxyphenyl methyl radical 77 shows similar dimerization to form a peroxy dimer with an interesting head-to-tail connectivity which is different from the tail-to-tail mode found for radical 1. With only four methoxy groups for steric blockage, radical 77 does not have enough steric protection against the penetration of oxygen to the central carbon so that it forms a peroxy dimer linked between a central carbon and the para carbon in that asymmetric third ring. We are in the process of determining the first step of the reaction by reaction of radical 77 in a hydrogen donating solvent; The preliminary results suggest that the first step is the attack of oxygen at the para carbon. The X-ray structure of the peroxy dimer (J acobson's ) 91 from radical 77 is shown in stereo view in Figure 45. Figure 45. Stereo View of X-Ray structure of head-to-tail peroxydimer 91. 171 3. Magnetic Behavior of Radicals and Complexes We believe that we have made magnetic materials in polycrystalline form from 1 and various lithium salts. I studied these blue-black powders by standard methods (SQUID, EPR) for magnetic susceptibility determination as a function of temperature and magnetic field strength. In our experiments, ferromagnetic behavior165 was reproducibly observed in the blue-black powders obtained when solutions of 1 and LiBF4 were mixed and evaporated to dryness under an argon atmosphere. Studies of magnetic field dependence in these materials showed hysteresis (Figure 46) a classic signature of bulk magnetism with far less spin counts than theoretically expected. The tetrafluoroborate anion was replaced by the similar-sized perchlorate ions without qualitatively affecting the field dependent magnetic properties. This result indicated that the bulk behavior is not due to chemistry involving the counterion. Radical 1 alone, a red powder, showed weak antiferromagnetism, in contrast to the salt-treated samples; lithium salts alone showed diamagnetic behavior, suggesting that the observed ferromagnetism must somehow derive from the combination of the radical with the lithium salt. 172 6O 40" 20" M(emu/mol) 0 q 1-LiBF4 -20 ‘ -40 - ‘60 j I ' l I 1 I I -6000 4000 -2000 O 2000 4000 6000 H (G) Figure 46. Plot of magnetization M (in EMU) vs. magnetic field strength H (in Gauss) at 5K for Radical 1 alone and for its complex with LiBF4. Iron from the laboratory environment is a frequent ferromagnetic contaminant in studies of magnetic materials. In light of the reproducible SQUID results above and our fear that contamination with iron or other ferrous metals might be causing the observed hysteresis, we devised a metal-free synthesis of radical 1. Martin's original synthesis used Cr(II) ion as the reducing agent; we showed that zinc works well too. However, we sought to avoid metals, such as Cr or Zn, that could carry ferrous contaminants. Ultimately, we discovered that ascorbic acid (vitamin C) and Iodide salts are both able to reduce salts of the cation in THF; the neutral radical then migrates into a nonpolar organic solvent layer, where it is easily separated and isolated. Various attempts were made to eliminate 173 potential iron contamination in our SQUID samples; careful filtration of radical and salt solutions before mixing, use of glass distilled solvents and high purity salts, chromatography of the starting radical solution over silica gel, etc. A series of SQUID samples were analyzed by atomic emission photometry which showed some iron at the level of the background, regardless of their SQUID behavior in all samples and controls. However, the levels found would be more than sufficient to yield the ferromagnetism observed, if the iron was in a ferromagnetic form. 3.1. Diamagnetism of Triaryl Methane 73 and Metal Salts The Pascal constants provide an empirical method for estimation of diamagnetic corrections. For the complex 1-LiBF4 the diamagnetic contribution is calculated to be -2.89 x 10'4 emu/mol. Greater accuracy can be obtained by the direct measurement of the susceptibility of a diamagnetic analogue of the paramagnetic compound which is of interest. For example, a diamagnetic correction for radical 1 can be achieved more accurately by a subtraction of magnetic measurement data from the corresponding triaryl methane 73 discussed earlier (section 2.4 of chapter I) which has one more hydrogen than the radical. Furthermore, by this process of diamagnetic correction we were able to eliminate the error factors inherent in the SQUID measurement which come from the method of sample handling, etc. The observed diamagnetic correction for radical 1 using the methane 73 gave similar diamagnetic contributions at the high temerature region. Figure 47 shows a typical field dependence of methane 73 at 2K showing characteristic diamagnetic behavior. 174 4-1 M (emu/moi) 0'l -2-1 ‘6 ' I ' 1 I fl I -6000 -40 -2000 0 2000 4000 6000 H (G) f I '7 1)) 0.4 0.21 M (emu/mol) 0.0- -0.2 " -0'4 —I 1 f If If r r7 I I ' -6000 -4000 -2000 0 2000 4000 6000 H (G) Figure 47. Field dependence of a) methane 73 at 2K and b) CdC12 at 5K. The induced molar magnetic moments decreases as the field strength increases which is the characteristic field dependence (inversely proportional) of a diamagnetic material. 175 3.2. Paramagnetism of Radicals and Radical Complexes As I discussed in the introduction, a paramagnet concentrates the lines of force provided by an applied magnetic field and thereby moves into regions of higher field strength. Paramagnetic susceptibility is generally independent of the field strength, but shows temperature dependence according to the Curie Law. Radical 1 shows a paramagnetic temperature dependence over a wide temperature range; it is weakly antiferromagnetic at low temperature, although the coupling is not a big one. We can show the paramagnetic properties of the radical in two different ways using the plot of 1/x vs T and xT vs T. Figure 48 show such plots with a slight antiferromagnetic downturn at low temperature. Most of HMTP methyl radicals behave similarly to radical 1 in their temperature dependencies. 176 1000 800 ' 600' 1/X (emu/mo!) 400' 200 " 100 200 300 Temperature (K) 5' OT‘ 0.4 0.3 ‘ xT(emu/mol) 0,2 J 0.1 " 0.0 - . fi —r f . 0 100 200 300 Temperatue (K) Figure 48. The temperature dependence of radical 1 analyzed in a) l/x vs T and b) xT vs T at 5000G with Curie constant of 0.34 emu/mol which corresponds to 91% of ideal S=l/2 radical. Presence of the diamagnetic peroxy dimer 91 reduces the total spin count. _ 177 The radicals 86 which I discussed earlier (section 3.2 in chapter I) were synthesized to build an analogue of 1 which could not form extended chains by pairing around metal cations. Radical 86 offers a bound Li+ ion a full coordination sphere of six ether oxygens, effectively "capping" the radical on both faces. Thus, the radical center in 86-2LiBF4 is encapsulated and should behave as a completely isolated paramagnet. \o \o‘ [O (0.99 (Q 3: 2 LiBFg (Q 3 . 2 BF; 0 0‘ ea 9.. 86 86°2LiBF4 As shown by CV data earlier, the radicals 1 and 86 are respectable reductants with nearly identical redox potentials. Fears about impurities in our studies of radical 1 with salts led us to speculate that even though no ferromagnetism was found in our starting materials, something ferromagnetic might be generated by reduction of a diamagnetic impurity. Radical 86 should be just as likely to effect the hypothetical reduction as 1, but magnetic measurements shows that the complex 86-2LiBF4 is a textbook paramagnet. In fact, these samples are among our best instances of isolated S=1/2 radical behavior. Thus, blocking chain formation in 1 significantly alters the nature of the LiBF4 complex products, consistent with the notion that such chains or at least the open faces are important to the complexation chemistry of radical 1. We have verified this 178 expectation by magnetic susceptibility studies of powders of the complex. Figure 49 shows plots of magnetic susceptibilities as a function of applied magnetic field and temperature. As expected for an isolated radical, a Curie plot (l/x vs. T) shows ideal paramagnetic behavior with a ueff of 1.56 1.113. We attribute this slightly low value to uncertainty in the exact radicalzLiBF4 ratio, and to a small amount of radical oxidation during preparation. There is no indication that adding LiBF4 to 86 puts the radical centers into communication. This finding is in contrast to the situation with 1. Like 1, radical 86 shows some tendency toward oxidation on treatment with LiBF4 under the air. This behavior is evidenced by the appearance of the blue color of the corresponding triarylrnethyl cation in each cases. This result reinforces the notion that the ferromagnetism seen from a lithium complex of radical l is unique to a radical which can bind metals on both open faces to form chains, and is not simply due to an artifact such as iron particle production by reduction of adventitious iron salts. 179 400 200 7 M (emu/mol) 0 ' -200 ‘ -400 I I I r I I F I I I -6000 -4000 -2000 0 2000 4000 6000 H (G) r b) 1000 800 ‘ 600" l/x 400 ‘ 200‘ 0 a T 4 a 4 0 100 200 300 TEMP (K) Figure 49: (a) Plot of M vs. H for 86-2LiBF4 at 5K; (b) Plot of 1/x vs. T for 86-2LiBF4 at 5000G showing paramagnetic behavior. 180 In the course of the NMR ion-binding studies, we developed a more detailed understanding of one counterion's role as a partner and ancillary ligand to Li+ bound in the tripod ether's pocket. These insights have translated into an explanation of the simple S=1/2 paramagnetic behavior found in magnetic susceptibility studies of 1-2LiI complexes. If the radical is capped by Lil subunits, it is sensible that it should behave as an isolated paramagnetic center like complex 86c2LiBF4. Figure 50 shows paramagnetic behavior of complex 1-2LiI at 5000G. 1.50 1.00 - ueff ( 11B ) 0.50 ‘ 0.00 . . ‘ ' ' ' 0 100 200 300 Temperature (K) Figure 50. Temperature dependence of complex l-2LiI at 5000G showing simple paramagnetism of the complex (81% of spins are counted). 181 By the same token, if the radical is capped by NH4I subunits, complex l-N H41 should behave as an isolated paramagnetic center like the case above. The temperature dependence of this complex is shown in Figure 51. l/x O 100 200 300 Temperature (K) Figure 51. Temperature dependence of complex l-NH4I at 500G showing simple paramagnetism of the complex by effective capping of the linear chain formation. At very high external fields and very low temperatures, the magnetization M becomes independent of field and temperature, and approaches the maximum or saturation magnetization Msat which the spin system can exhibit. This situation corresponds to the complete alignment of magnetic dipoles by the field (Msat=5.58 x 103 erg/Os mol for g=2, S=1/2). 182 8 7 b : ...............U I .. 6 _- .0. O AAAAAAA‘9446AA” 5 - o. .e‘“ C A‘ +++++++++ : A ++++++ M/NuB 4 '. ' . E 0 A +++ .0000000009"'” I A + .... 3 r 04 + .0. 000000000 ' i ° 00000000 A O 2 -o * ° 0000 ...o :°+o 00 °°°°°°°°°°°° 1:-‘”o° 000 ......... °°°' Ito.° °°°°°°°° '§:2°l°.. l l l o l llllllllllllllllllll 0 05 1 1.5 2 25 3 71 Figure 52. Plot of magnetization per radical, expressed in Bohr magnetons, against 11 for S = 1/2 to 7/2 (bottom to top). It is convenient to calculate the magnetic moment of a magnetic system with arbitrary spin-quantum number. The average magnetic moment of one radical can be calculated using the Brillouin function B301) which is defined below for N non-interacting atoms. Msat = ngBSBS(n) 8113 E; n=kT 1 1 1 1 35(1)) = —§[(S+ -2—)coth(S + -2-)n — Ecom—g] N is Avogadro's number, g is Landé's constant, S is the spin-quantum number. 183 The mean magnetic moment , or magnetization can be written as below. A plot of magnetization per radical expressed in Bohr magnetons 11B, against 1] (dimensionless parameter directly proportional to fieldstrenth) for S: 1/2 to 7/2 is shown in Figure 52. < #2 >= gflBSBSU?) M = N < 0, >= NgrtBSBsW) Most of our radicals do not show saturation magnetization even at 55000G at 1.8K which are the field and temperature limits of our SQUID at Michigan State University. However in the case of the complex l-LiI, a near perfect fit is found to the Brillouin function. (Figure 53 ). Many of our radicals might show saturation at higher than 55000G using a pulsed magnet. In some of the organic radicals such as the DPPH complex with benzene which forms a linear chain in the solid state, Gerven used a pulsed magnetic field to achieve saturation of the complex at 13.4 T, more than twice of the field strength of our SQUID! 166 184 M/NuB Figure 53. The saturation magnetization of the complex l-LiI at 1.8K. The data were fitted to Brillouin function using Kaleidagraph 3.0 with S=0.62 and R=0.9999. 3.3. Ferromagnetism of Complexes Ferromagnetic behavior was observed in the precipitates obtained when solutions of 1 and LiBF4 were mixed under argon atmosphere. Radical 1 alone, a red crystal, showed a weak anti ferromagnetism, in contrast to the salt-treated samples; lithium salts alone showed diamagnetic behavior, suggesting that the observed ferromagnetism must somehow derive from the combination of the radical 1 with the lithium salts. Along the same line, the complex of radical 1 with ZnC12 shows a hysteresis curve with less saturation than complex l-LiBF4 as shown in Figure 54. 185 Our results are reproducible despite variations in synthesis of radical 1, careful control experiments on starting materials, and filtration to remove any magnetic particulates. Atomic emission experiments yielded inconclusive results for the presence of ferromagnetic impurities. An ac susceptibility measurement, which is now available with the new SQUID at Michigan State University might also give better handles when we have structurally well defined materials. 186 M (emu/mol) 0 4 / 1 / -3000 ——v -1500 b) 0 H (G) 1500 3000 15 10' 5- M (emu/mol) 0- -5- —10 ‘ -15 -5500 f—I -1500 f m -3500 500 H (G) f —v 2500 4500 Figure 54. The field dependence of a) the complex l-LiBF4 at 5K and b) the complex loZnClz showing characteristic ferromagnetic hysteresis curve. 187 3.4. Antiferromagnetism of Radical Complexes According to our original scheme for building magnetic chains, control over magnetic properties can be achieved by varying the metal cation, and thus, the distance between radical centers. The first consideration in selecting metal salts was the size of the metal cation and the availability of unpaired electrons in the metal. The complex of triaryl methyl radical and a paramagnetic metal cation might give us better model to study the linear combination of hetero Open shell environment. Various alkali metal salts were the logical choice at first and some first row transition metal salts like CuC12 and CrCl3 were tested for possible magnetic chain formation. None of the sodium or magnesium complexes with radical 1 showed any interesting magnetic behavior except simple paramagnetism which suggests that the sodium cation is unable to induce a magnetic interaction to occur. The complex of l-CrCl3 again showed simple paramagnetic behavior with far less spin count than expected. The significance of this observation is not clear. Perhaps the metal caps the radical center as in the case of l-LiI and loN H41 and there are antiferromagnetic interactions between the unpaired electrons of the metal and the radical. It is not clear what kind of interactions are involved without knowing the solid state structure of the complex. The complex 1-CdC12 shows classical antiferromagnetic temperature dependencies with xmax at 25K which may be explained with an alternating linear chain Heisenberg model. Attempts were made to fit the temperature dependence data with various models. The coupling constant J varies from -10.2 to -19.7 cm-1 depending on the model used (Figures 55 and 56). The Bleaney-Bowers dimer model with paramagnetic impurity correction gave an inter-radical coupling constant of J = -l3.5 cm-1 (-19.4 K) with 5% paramagnetic impurity (R =0.9998). Calrin‘s mean field correction of the Bleaney-Bowers model (described in the introductions) with paramagnetic impurity correction gave J = -13.6 cm'1 (-19.5 K), J'Z= 11.4 cm“1 (16.5 K) where Z is number of neighboring dimers with 5.6% paramagnetic impurity (R = 0.9998). This alternating dimer model (chain of dimers) can be explained as an alternating Heisenberg linear chain when Z is 2 with an interdimer coupling constant of J ’= 5.7 cm-1 (8.3 K). It is still not clear which model is the best for the complex l-CdC12 without having the solid state structure of the complex. The anti-ferromagnetic behavior of the complex l-CdC12 was reproduced several times; this combination is unique in the sense that the complexes of radical 1 with CdBr2 and Cdlz only show simple paramagnetic temperature dependencies, as do those with NaBPh4. Numerous attempts to grow a single crystal were unsuccessful, with only polycrystalline materials Obtained. X-ray powder patterns (Figure 57) of the radical and the CdClz complex show clear differences but do not provide real structural information for this unique system. 189 0.03 “emu/mol) 0.02 1 0.01 r-Cdcn 0 50 100 150 200 250 300 350 ,7, Temperature (K) ' 0.01 0.008 0.006 X (emu/mol) 0.004 O [JUIJULIUUIJULJILHLILILIIIU 0 50 100 150 200 250 300 350 Temperature (K) Figure 55. The temperature dependence of a) the radical 1 and complex 1-CdC12 at 500G; b) with Bleaney-Bowers dimer model with J = -13.5 cm'1 (-19.4 K), 5% paramagnetic impurity, and R ==0.9998. 190 0.01 0.008 0.006 X (emu/mol) 0.004 0.002 0 LUIJJIJJIJJIILIILIIIJIIJWLIJJU O 50 100 150 200 250 300 350 Temperature (K) 0.008 - 000 0.006 x (emu/mol) 0.004 0.002 b 0 j[JIJLI”DULJIJIHILILILILILILIHI 0 50 100 150 200 250 300 350 Temperature (K) Figure 56. The temperature dependence of the complex 1-CdC12 at 500G a) with alternating dimer model with J = -13.6 cm'1 (-l9.5 K), J'= 3.8 cm-l (5.5K), Z=3, 5.6% paramagnetic impurity, and R = 0.9998; b) with linear chain Heisenberg model J = -19.7 cm:1 (- 28.5K) and R = 0.9884. 191 1800 % El) o—«v 1‘ 1200 600 ..—...——.-._..~_.—._’-. -— 5 1o 15 20 25 30 2 THETA RISAKU / USA 8000 - ' 1 I 4000 4 3 4 7 'o e . 6 ' 9 5 __ 4 AA A A A A‘ o fiff vrjri—va—vy—v—Vfifivvyvv —v rvIfifiv —rfw V—fj 5 10 15 20 25 30 2 THETA HIGAKU / USA 4 1500 * C) '. 2 1200 + s 600 o t—‘raT—vfifi—va—‘~Vfifv—r’f. .yv .‘ .v v v 1 5 10 15 20 25 30 2 THETA RIGAKU / USA Figure 57. X-ray powder patterns of a)CdC12, b) the radical 1, and c) the complex l-CdClz. 192 An attempt was made to find information on the electronic coupling of Cd(II) and the radical center by ESEEM using radical 86. Although significant new features were observed, the spectra were too complex to yield any meaningful answers. The experiment is complicated by the fact that there are two different magnetic Cd isotopes (111 and 113), roughly equal in natural abundance, and a third non-magnetic isotope (112) that makes up the majority (~75%) of the element's composition. It would be very useful to have the structure of the complex in order to understand the mechanism of the magnetic coupling between CdC12 and 1. There are two possible coupling routes to form an alternating dimeric chain: a) the originally intended linear chain formation through metal ion-binding oxygen pockets to form a chain; b) simple Cd-mediated linkage from para- carbon to para-carbon in two radicals as shown below; or both. The magnetic data from CdBr2 and Cd12 suggest that the mechanism shown below may not be the right one. The magnetic measurements on complexes of other para substituted radicals such as 78 may give more straightforward answers to this question. \ / 0 o f“ $1 @o O ‘1“ 0 C1 ()0 l \ 4. Design of Molecular Magnets by Other Ion-Binding Recently, we found that treatment of TEMPO with ammonium or lithium salts leads to high-spin coupled materials. EPR spectra for these 193 systems Show strong half-field transitions, diagnostic for high spin- coupling. No structural details are yet available, but the Lil complex of TEMPO forms crystalline long needles, and we are optimistic that it might provide a probe for the long overdue question of pairwise radical coupling directly through alkali metal cations. < N—OIIIM+IllO—N > < N—O”""H— +-HIII||IO'—N > It is tempting to envision pairwise radical interactions through a metal unnflj—Z—munu cation in the case of TEMPO and DPPH. In contrast, the resonance delocalized structures of galvinoxyl and nitronyl nitroxides suggest that these subunits may be able to participate in forming extend structures. Special attention should be paid to the search for triplet or higher spin transitions in the EPR spectra and susceptibility measurements using SQUID. + O -O--M+- O-N O--M+~ H H £1 ----- H—N+ ---H O~<::E\I-O-m-H--N+ H"- -0- $70.12? -H—N+ -H----- Furthermore, other commercially available stable organic radicals such as TEMPOL can be used to probe this problem including the effects of the counter anions of the coupling salts or the formation of extended structures such as those shown above. 195 5. Introduction of Cross-Linkers To obtain materials which exhibit bulk ferromagnetism, high-spin coupling of the constituent paramagnets must occur in three dimensions. Our picture of stacked radicals sandwiched around metal ions provides a one-dimensional coupling scheme, but how can the expected stacks be made to communicate magnetically ? In the few known molecular ferromagnets, this issue has simply resolved itself without explicit action being taken to control it. However, to understand the requirements for stack-stack coupling in detail, we planned to examine both intra- and interstack types of electron-electron coupling by explicitly building biradical systems to model and/or induce these interactions. By isolating the functional diradical unit 2 of our proposed polymeric chains, the intrastack electronic coupling can be studied. We planned to build two series of these diradical complexants; the triarylrnethyl radicals linked through alkyl chains and meta-xylylene linked poly radicals. An attempt to build the alkyl chain linked biradical shown below using monomer 92 was unsuccessful. "(OOHO cho3 __ 2 6 BrMBr HO 1 196 To avoid chain-chain cancellation of magnetic moments, an interchain coupling mechanism is necessary. An approach to this requirement is to build electronically and physically cross linked monomers into the chains, enforcing interchain communication. In contrast to 92, aromatic ring linked diradicals should be fairly rigid, with little opportunity to vary the radical-radical distance, whether or not a metal ion is present in the internal cavity. There are many possible model systems which could be used in building three dimensional structures. Among numerous possible aromatic linked biradicals, several phenyl linked ones, such as diradicals 93 and 94 are synthetically accessible by simple modification of usual procedure. According to the n-topology of altemant hydrocarbon discussed in the introduction, the diradical 93 should be high-spin coupled and the diradical 94 should be low-spin coupled. Radical 93 shows two reversible redox waves in the cation/radical couple of -O.12 and -O.41V, and only one redox wave in the 197 anion/radical couple of -2.03 V in CV (Referenced to ferrocene oxidation; Measured in CH2C12 solution of (n-Bu)4NBF4 ). The two reversible redox waves in the diradical 93 presumably correspond to the radical cation and the diradical which is characteristic evidence for the formation of a diradical. The CW-EPR spectrum of the diradical 93 in THF was too complex to analyze for hyperfine coupling constants at present.167 The half-field transition of diradical 93 was also not detected under various conditions. Ar Ar ° ° Ar Ar Ar Ar Ar Ar 93 (Ar=2,6-dimethoxyphenyl) 9 4 A diradical such as 95 could also be used in building three- dimensional structures with proper ion-binding ability. This system shows a reversible redox potential of -O.44 V for the cation/radical couple and -1.93 V for the anion/radical couple (Referenced to ferrocene oxidation; Measured in CH2C12 solution of (n-Bu)4NBF4 ), and a CW-EPR spectrum which is comparable to those of other para substituted TMTP methyl radicals such as 78, 79, 80, and 83 which means this radical is not antiferromagnetically coupled strongly as predicted by theory. A half-field transition was not detected. 95 (Ar=2,6-dimethoxyphenyl) 198 A triradical such as 96 could also be used in building extended structures by using two-dimensional frameworks extended into three- dimensional structures by ion-binding. Characterization of such systems is complex and the already detected half—field transition spectra of 96 at 77K has a problem of interpretation due to the fact that it might also be the half- field transition of the diradical cation or a diradical as shown below. 199 CONCLUSIONS AND SUGGESTIONS FOR FUTURE WORK Conclusions We have begun a new line of investigation into molecular magnetic materials, which may ultimately extend our understanding of molecular magnetism. Magnetic measurements (SQUID) suggest that we can induce ferro~ and antiferro-magnetic behavior in complexes of stable organic radicals with various metal salts. Cyclic voltammetry of HMTP methyl and TMTP methyl radicals demonstrates our ability to adjust their electronic character by variation of substituents. X—ray data for radicals 1, triarylrnethyl cation 62, triarylborane 63, triaryl carbinol 71, triaryl methane 73, triarylsilane 74, and ammonium adduct 72, and the hexachloro ammonium adduct 75 have built a structural basis for preorganized ion-binding. Complexation studies of 63, 64, 87 , and 88 by UV—Vis spectrophotometry and NMR suggest complexations. ESEEM studies of the truncated complex 86-2LiBF4, 86-2LiI, 86~2NaBPh4 , and 8602NaI show that two metal ions are bound at the expected distance with spin densities of 0.3% on the metal cations, and molecular mechanics and semiempirical molecular orbital calculations agree on the structure of the complex. Ab Initio molecular orbital calculations offer an explanation for the high-spin coupling in radical metal radical systems. Two critical issues have focused our efforts in this research. First, structural characterization of metal-radical complexes is essential to our understanding of the rules governing long-range organization in these new materials. Such an understanding is required for the rational construction 200 of organic magnetic substances. By using X-ray crystallography, we have studied monomers with the binding sites of interest; we have also obtained a certain amount of indirect (spectroscopic, electrochemical) evidence for ion binding. Now more detailed geometrical insights into the complexation phenomena are needed. The second focal issue is the fundamental description of the metal ion-mediated magnetic interactions between organic paramagnets. Assuming our structural pictures of the extended complexation are correct, how is a lone radical such as 1 perturbed by metal ion binding? More importantly, how do two radicals 1, magnetically independent in the absence of salts, couple around a metal ion to make an interrupted G-bond? What is the sign and magnitude of the exchange integral between the two paramagnetic centers? Can it be properly understood in terms of simple molecular orbital pictures? We have begun with magnetic, chemical, and structural characterization of various triaryl methyl radicals and their complexes with different metal cations. There is enormous structural flexibility in our ion- binding approach to building molecular magnet, where the triaryl-X platforms provide triether binding pockets for metal ions. There are several controllable variables in the basic triaryl-X structure, and in terms of the number of triether binding faces present. All of these variables should be further investigated in order to optimize the selection of the best candidates for assembly of stable extended complexes. X-ray structures for complexes of radicals and boranes with metal salts are desperately needed. Detailed magnetic characterization of well-defined materials, ideally as single crystals, will enhance our understanding of the 201 relationships of structure and odd—electron coupling. By systematically building up an understanding of the ion binding and magnetic properties of these systems, we hope to establish this new approach as an effective paradigm for the assembly of 1— and ultimately of 3-dimensional molecular magnetic materials. Suggestions for future work 1. Aromatic Linked Biradicals We planned to attach a second hexamethoxytriphenyhnethyl radical unit at the para position of an aryl ring of 1 to make diradical 97. Studies on Chichibabin's hydrocarbon168 suggest that for the highly twisted 97 , open-shell paramagnetic electronic structures will be important, despite the availability of closed-shell classical structures.169 In hopes of extending our studies to the rational assembly of metal-organic complex solids with bulk magnetism, the examination of the complexation and chain-forming behavior of 97 , 98, and their derivatives will be very important. 202 98 2. Magnetic Coupling through Hydrogen Bonds Numerous examples of macromolecular chemistry have been based on frameworks of hydrogen-bonds.”0 Extensive studies have been reported of the structural and electronic properties of such systems. The detailed descriptions of orbital interactions through hydrogen-bonds lead us to consider a strategy in building organic magnets. Electron transfer and energy transfer through space and through bonds, especially through hydrogen bonding have been pursued extensively in recent years, and theoretical understanding of the interactions involved can be explained by many theories.”1 Magnetic interactions through bonds and through space have been proposed by Hoffmann and Gleiter in organic diradicals and in inorganic bimetallic bridging ligand systems and the basic theory has been accepted for more than 20 years.172 Magnetic interactions through hydrogen-bonds were suggested by Carlin in 1985 in an attempt to explain magnetic susceptibility data on a 2:1 complex of nitroxyl radical and Cu (11) complexes. In an attempt to explain the structural and magnetic behavior of a TCNQ '8' 203 complex of radical cation of 1,4-hydrazine systems shown below, Koch reported at the 1993 Denver ACS meeting the first example of magnetic interactions in an infinite organic chains in which the ferromagnetic coupling was mediated by hydrogen-bonding interactions. The possibilities of designing magnetic interaction through hydrogen bonding are endless in organic radicals as illustrated by the numerous macromolecular structures already reported. Any stable organic radicals with proper functions for hydrogen bonding can be considered to be a target for testing the proposed magnetic coupling. Simple derivatives of TMTP radical such as radicals 82, 99, and 100 shown below (which we have synthesized already) can serve as pairwise model systems in building organic magnets. 204 3. Ion-Binding in Various Nitroxyl Radicals As discussed in the introduction, the study of nitroxyl radicals is one of the most active areas of molecular magnet research. The nitroxyl radicals, especially diaryl nitroxyl radicals, offer logical extensions of our strategy in building molecular magnets by ion-binding. The 4,4'- dimethoxy diphenyl nitroxide shown below has never been studied for molecular magnets even though the X-ray structure with a packing diagram has been known for more than 40 years. Magnetic measurements on a 205 single crystals of the radical might give insight on the mechanism of inter- radical interactions in nitroxyl radicals. O,© :O There are several advantages of using nitroxyl radicals compared to triaryl methyl radicals: The Lewis basicity of nitroxyl as demonstrated by Drago in TEMPO; spin polarizability of nitroxyls; inherent stability of diaryl nitroxyls when they are protected against CO dimerizations by para- substitution and twisting of rings by ortho-substituents. Possible linear complexes of bis-(2,6-dimethoxyphenyl nitroxyl and bis(2,6-pyrimidyl) nitroxyl with metal ions are shown on the next page. \0 (I) O/ (I). N N\ N N\ a?) or To 206 As we have seen earlier, the para-pyridyl nitronyl nitroxide is ferromagnetic in the solid state and forms linear chains through "hydrogen-bonding" interactions. The para-pyridinium nitronyl nitroxide might crystallize to give a magnetic structure with a hydrogen~bonding framework shown below. There would be little overlap in this arrangement between the SOMOs, while maintaining overlaps between other frontier orbitals which is one of the requirements for ferromagnetic coupling. \ N+ H'm "0° °’N /+~N~O ..... ).>—<3 Iz‘i'I-m". "0’ I 2/ 2-0m 'I-Z+ \ """ Of‘N-F’ N'O."" 'H— N+_ \ ;élt 207 Ullman's biradica1130 is another logical precursor for an extension of our non-covalent interaction strategy. A prOposed Ullman's biradical complex with NH4+ is shown below as a stereo view. There will be little overlap in this diradical between the SOMOs because of the orthogonality of the orbitals, while maintaining overlaps between other frontier orbitals. H H N+ I \ 1-1 H 0* 0' I \ N+ N H \ N N“ \ I Q' 9' H\ .H N4- H H' : ~ 208 EXPERIMENTAL 1. General Procedures All 1H NMR and 13C NMR spectra were obtained with a 300MHz Varian Gemini or a 300 MHz Varian VXR-300 instrument. IR spectra were recorded using Perkin-Elmer 599 IR or Nicolet IR/42 spectrometers. UV Spectra were recorded on a Shimadzu UV-l60 or Hitachi U-2000 spectro- photometers. Gas Chromatography was conducted on a Perkin-Elmer 8500 using a 25-m (0.32 mm Diameter) bonded methyl silicone column with an FID detector. HPLC was conducted on a Perkin-Elmer Binary LC Pump 250 and LC-235 Diode Array Detector. Mass spectra were recorded on a Finnigan 4000 GC/MS or on a VG Trio-1 5890 GC/MS, using a 30m SE-54 Altech Capillary column. Electron Spin Echo Envelope Modulation (ESEEM) studies by pulsed EPR esperiments were conducted on an instrument build by Professor John McCracken, MSU. Powder X-ray pattern were recorded on a Phillips XRG 3000 computer— controlled powder diffractometer, operating at 40kV, 35 mA. Single crystal X-ray diffraction peaks were recorded on a Nicolet P3F diffractometer or Enraf-Nonius CAD-4 diffractometer using graphite monochromated Cu Ka radiation and an w-20 scan by Professor Kahr at Purdue University. Absorption corrections were applied using the program DIFABS (Walker & Stuart, 1983), SHELXS86 (Sheldrick, 1986), and MULTAN82 (Main, Fiske, Hull, Lessinger, Germain, Dec lercq, and 209 Woolfson, 1982). In each case computations were performed on a VAX computer using SDP/V AX software (Frenz, 1978). 2. Solvents and Chemicals All chemicals used were from Aldrich Chemical Co. or Lancaster Synthesis Inc., unless otherwise noted. Some of the metal salts were from Alpha Chemical Co. Solvents were from Fisher Scientific Co., MTM Research Chemicals, J. T. Baker Chemical Co., Johnson Matthey Chemical Co, and Mallinckrodt, Inc. THF, diethyl ether, and benzene were freshly distilled from sodium and benzophenone under argon. Dry acetonitrile and other solvents were distilled from CaH2 under argon as needed. 3. Equipments and Procedures CV Measurements CV spectra were recorded on a EG & G Princeton Applied Research Scanning Potentiostat Model 273 and 362 using Bioanalytical Systems Ag/AgCl as a reference electrode. A potentiostat Model 273 was interfaced to a PC which controls the potentiometer, including data collection and plotting. The construction of the experimental cell has been described elsewhere. (P. T. Kissinger, J. Chem. Educ. 60, 702, 1983.) Measurements were made under argon on a 1x 10*4 to 1x 10*5 M sample of substrate in 7 to 8 ml of freshly distilled dry THF or spectroscopic grade 210 methylene chloride containing 0.1 M tetra-n-butyl ammonium tetrafluoroborate. A platinum button was the working electrode and a platinum coil was the counter electrode. The reference electrode was a silver wire immersed in a 3M NaCl solution, separated from the bulk solution by a Vycor tip. The Vycor tip gave some water leakage problems, but it was satisfactory in determining redox potentials with ferrocene as an internal reference to determine the potentials. Pulsed differential voltammograms were obtained with a slight modification in operating programs in the PC. The scan rates were 50 to 400 mV per second and a typical one was 200 mV per second. In all cases reversibility was checked by verifying the equal height of the oxidation and reduction waves in cyclic voltammetry, and by examining peak to peak distances of the forward and I'CVCI‘ SC scans. VT-EPR Measurements CW-EPR spectra were recorded using a Varian E4 spectrometer with variable temperature control. The temperature range was 25 to —140 °C (temperature was controlled by a flow of nitrogen gas running through a cooling coils in liquid nitrogen). Most of the spectra for HMTP methyl derivatives and TMTP methyl derivatives gave strong enough signals without phasing problems, except in the range of -70 to -90 °C, which was evidently inherent to the spectrometers. Various conditions were studied with scan windows of 10G to 4000G, modulation amplitude of 1 x 10'2 G to 1G and receiver gain of l x 101 to 5 x 103 with microwave power of l to 5 mW. SQUID Measurements Magnetic susceptibility measurements in the temperature range from 1.8K to 350 K at -55kG to 55kG were performed on a MPMS Quantum Design SQUID magnetometer. Powder or polycrystalline samples were dried, weighed and placed into a capsule with a minimum amount of cotton on the top. A ventilating hole was made in the capsule with a fine needle, the capsule was placed in a normal drinking straw, and it's position was fixed securely with threads up and down. The sample straw was attached to a sample rod with general centering according to the length of the sample from the t0p of the rod with the mass of the sample, and inserted into the evacuation chamber. The sample was lowered into an experiment chamber slowly followed by three cosecutive evacuations of air. The sample was oriented in the 200 G to 1000 G external magnetic field according to the centering scheme built into the machine. Magnetic susceptibility measurements were conducted according to temperature and field dependence sequences. A general sequence consisted of field dependence measurements over -5000 G to 5000 G at 2 and 5 K, magnetization measurements from 0 G to 50000 G at 1.8 K and temperature dependence experiments from 0 to 320 K at 200, 500, and 5000 G field strengths. The data collected were extracted into a data file in IBM PC ASCII format consisting of temperatures in K, external field strength in gauss, induced magnetic moments in emu, and their standard deviations. The data files were converted as Macintosh format and analyzed and plotted as molar susceptibilities. L_— 212 4. Synthesis General Procedures for Generation of Radicals Method A: The cation tetrafluoroborate salts (200 mg) were reduced by 10 ml of ~5 M CrC12 solution (prepared with CrCl3-6H20 and Zn/Hg in 10% HCl solution), extracted with 10 m1 of methylene chloride, washed with water twice, and dried with magnesium sulfate. Flash column chromatography with diethyl ether over silica gel gave a solution of the radical. Method B: The cation tetrafluoroborate salts (200 mg) were reduced by the addition of an excess (5 equivalents) solid CrC12 into 10 ml of methylene chloride solution of the cation and water (1 :1 in volume) and the radical was extracted into methylene chloride, washed with water twice, and dried with magnesium sulfate. Flash column chromatography with diethyl ether over silica gel gave a chromatographically pure solution of radical. Method C: The cation tetrafluoroborate salts (100 mg) were reduced by the addition of an excess iodide salt (3 equivalents) (NaI, Lil, NH4I, and n- Bu4NI) solution in 20 ml of dry THF and the radical was flash chromatographed using diethyl ether over silica gel as in Method B. Method D: The cation tetrafluoroborate salt 4 (100 mg) was reduced to radical 1 by the addition of vitamin C (200 mg) in 20 ml of dry THF 213 followed by stirring overnight under argon at room temperature. The radical solution was flash chromatographed as above. Method E: Tris-(2,6-dimethoxy-3,5-dichlorophenyl)methyl ammonium tetrafluoroborate 17 was converted into a radical by diazotization. To a precooled solution of the salt (100-200 mg) in conc. HCl, a stoichiometric amount of saturated solution of NaN02 was added dr0pwise and the mixture was stirred until it warmed up to room temperature. The radical was extracted with 30 ml of diethyl ether, washed with water twice, dried with magnesium sulfate. Flash column chromatography using diethyl ether over silica gel gave a chromatographically pure solution of the radical. nrlPr d r frPrifiation f tion A solution of the triaryl carbinol (200 to 500 mg.) in diethyl ether was treated with a small amount (0.1 to 0.5 ml. ) of cone. HBF4, and the dark blue precipitate was filtered, washed with diethyl ether, and dried to give cation tetrafluoroborate salt as a crystalline solid. nrlPr r frPr rinof rinl 1/3 JUL 0 ; Method A: 1/2 JL / 0H Ar 0 Jr t Ar AIAI Method B: 0 Ar AT A Method C: 214 Method A: To a solution of 2,6-dimethoxyphenyl lithium (0.218 mol.) in 200 m1. of dry diethyl ether, 0.3 equivalent amount (0.720 mol) of dimethyl carbonate in 300 ml. of benzene was added dr0pwise (using a pressure equalized dropping funnel) under argon, and the mixture was refluxed for 3 days. The reaction mixture was poured into 600 ml. of water and the organic phase was concentrated. Crystallization from ether or methylene chloride in n-hexane usually gave crystalline carbinols. Method B: To a solution of 2,6-dimethoxyphenyl lithium (0.218 mol.) in 200 ml. of dry diethyl ether, 0.5 equivalent amounts (0.109 mol) of the esters in 400 ml. of benzene were added dropwise through a pressure equalized dropping funnel under argon and the mixture was refluxed for 3 days. The reaction mixture was poured into water and the organic phase was concentrated. Crystallization from ether or methylene chloride in n- hexane usually gave crystalline carbinols. Method C: To a solution of 2,6-dimethoxyphenyl lithium (0.05 mol.) in 60 ml. of dry diethyl ether, an equivalent amount of the ketone (0.05 mol) was added dropwise through a pressure equalized dropping funnel under argon and the mixture was refluxed for 3 days. The reaction mixture was poured into 100 ml of ice water and the organic phase was concentrated to yield a gray residue. Crystallization from ether or methylene chloride in n-hexane gave carbinols. 215 2,6,2',6',2",6"-Hexamethoxytrip_henyl methyl, HMTP 1: 2,6-dimethoxymethylbenzoate: 2,6—dimethoxybenzoylchloride (100 g.) in 1 liter of methanol with 5% 2,6-dimethoxybenzoic acid was refluxed overnight under argon with stirring. (Commercially available 2,6- dirnethoxybenzoyl chloride contains 5% 2,6-dimethoxybenzoic acid and it has been used as a catalyst for the reaction.) The solvent was evaporated and the residue was extracted with diethyl ether, washed with 5% NaHCO3 solution, dried with magnesium sulfate, and crystallized to give the theoretical amounts of ester as white crystals. 1H NMR (CDC13, 299.95 MHz): 3.80 (s, 6H), 3.89 (s, 3H), 6.54 (d, J=8 Hz, 2H), 7.26 (t, J=8 Hz, 1H) 13C NMR (75.43 MHz): 52.04, 55.58, 77.02, 103.43, 130.68, 156.86. di-(2,6-dimethoxyphenyl)methanone: A solution of phenyl lithium was prepared by the addition of bromobenzene (39.2g., 0.250 mol.) in 125 ml. of ether to lithium wire (4.0 g., 0.577 mol.) and 60 ml. of ether. Resorcinol dimethyl ether (30.0 g., 0.218 mole) was added and the reaction mixture was allowed to stand at room temperature under nitrogen for 60 hours. The reaction mixture was poured into 43.75 g. (218 mmol.) of 2,6- dimethoxy benzoilchloride in 400 ml. of cold diethyl ether and the reaction mixture was stirred for 3 hours at -78°C. The reaction mixture was poured into 300 ml. of ice water and the organic phase was dried and concentrated to yield a gray residue. Recrystallization from ether gave 25 g. of the product (38%). 1H NMR (CDCl3, 299.95 MHz): 3.67 (s, 12H), 3.50 (d, J=8 Hz, 4H), 7.21 (t, J=8 Hz, 2H) 13C NMR (75.43 MHz): 56, 104, 111, 131, 158 m/z 302 (M+), found 302. 216 Tris-(2.6-dimethoxyphenyl)methyl carbinol 71: A solution of phenyl lithium was prepared by the addition of bromobenzene (39.2g., 0.250 mole) in 125 ml. of ether to lithium wire (4.0 g., 0.577 mol..) and 60 ml. of ether. Resorcinol dimethyl ether (30.0 g., 0.218 mole) was added and the reaction mixture was allowed to stand at room temperature under nitrogen for 60 hours. Dimethyl carbonate (6.49 g., 0.072 mole) in 400 ml. of benzene was added and the reaction mixture was refluxed for 3 days under nitrogen. The reaction mixture was poured into 600 ml. of water and the organic phase was concentrated to yield a gray residue. Recrystallization from ether gave 22.4 g. of the carbinol. (71 %) 1H N MR (CDCl3, 299.95 MHz): 3.42 (s, 18H), 6.46 (d, J=8 Hz, 6H), 6.82 (s, 1H), 7.02 (t, J=8 Hz, 3H) 13C NMR (75.43 MHz): 56.21, 78.22, 105.98, 126.11, 126.85, 158.49, m/z 440 (M+), found 440. A solution of the radical in diethyl ether was prepared according to general Method A. Tris-(2,6-dimethoxyphenyl) borane 63: A solution of phenyl lithium was prepared by the addition of bromobenzene (39.2 g., 0.250 mole) in 125 ml. of ether to lithium wire (4.0 g., 0.577 mol..) and 60 ml. of ether. Resorcinol dimethyl ether (30.0 g., 0.218 mole) was added and the reaction mixture was allowed to stand at room temperature under nitrogen for 60 hours. 10.22 g. (8.85 ml., 72 mmol.) of BF3-OEt2 in 300 ml. of benzene was added and the reaction mixture was refluxed for 3 days under nitrogen. The reaction mixture was poured into 400 ml. of water and the organic phase was concentrated to yield a gray residue. Recrystallization from diethyl ether gave 11 g. of the borane (36%). 1H NMR (CDCl3, 217 299.95 MHz): 3.43 (s, 18H), 6.46 (d, J=8 Hz, 6H), 7.16 (t, J=8 Hz, 3H) 13C NMR (75.43 MHz): 56.52, 105.09, 130.17, 162.54 m/z 422 (M+), found 422. Tris-(2.6-dimethoxynhenyl-3.5-dichlorophenyl) borane 64: To a precooled solution of 3.7 g. (8.77 mmol.) of tris-(2,6-dimethoxyphenyl) borane in dry diethyl ether (to act as a base) and methylene chloride solution (10 ml. ether, 30 ml. methylene chloride), 4.5 ml. (7.1 g., 55 mmol.) of sulfurylchloride was slowly added through syringe at -78 oC and refluxed overnight with stirring under argon. The reaction mixture was washed with water twice, dried with magnesium sulfate, concentrated, and chromatographed with 10% ethyl acetate in n-hexane to give white crystals (3.0 g., 54%). 1H NMR (CDCl3, 299.95 MHz): 3.42 (s, 18H), 7.42 (s, 3H) 13C NMR (75.43 MHz): 61.09, 129.87, 133.00, 156.20 m/z 629 (M+), found 629. Tris-(4-chloro—2,6-dimethoxyphenyl)methanol 66: To a cooled solution of 5.0 g. (29 mmol.) of 3-chloro-dimethoxybenzene in 50ml of diethyl ether, 11.6 ml. (29 mmol.) of 2.5M n-BuLi in n-hexane was added slowly at -78°C under argon with stirring. The mixture was stirred at the room temperature for 2 days, 862 mg. (9.6 mmol.) of dimethyl carbonate in 50 m1. of benzene was added, and the reaction mixture was refluxed for 3 days under nitrogen. The reaction mixture was poured into 100 ml. of water and the organic layer was separated, dried with sodium sulfate, and the solvent was evaporated (200 mg., 4%). Cation: 1H NMR (CDC13, 299.95 MHz): 3.64 (s, 18H), 6.47 (s, 6H) 13C NMR (75.43 MHz): 56.90, 77.05, 105.63, 149.16, 162.30. 218 Tris-(2.4,6-trimethoxyethoxyphenyl)methanol 67: To a cooled solution of 20 g. (119 mmol.) of trimethoxybenzene in 100ml of diethyl ether, 48.4 ml. (121 mmol.) of 2.5 M n-BuLi in n-hexane was added slowly at -78°C under argon with stirring. The mixture was stirred at the room temperature for 2 days, 3.57g (39.6 mmol.) of dimethyl carbonate in 200 ml. of benzene was added, and the reaction mixture was refluxed for 3 days under nitrogen. The reaction mixture was poured into 300 ml. of water and the organic layer was separated, dried with sodium sulfate, and the solvent was evaporated (13.4g, 64%). 1H NMR (CDC13, 299.95 MHz): 3.44 (s, 18H), 3.75 (s, 9H), 6.04 (s, 6H), 6.66 (s, 1H). 13C NMR (75.43 MHz): 54.65, 77.02, 91.94, 157.77, 158.66. Cation: 1H NMR (CDCI3, 299.95 MHz): 3.57 (s, 18H) 3.97 (s, 9H) 6.03 (s, 6H). Tris-12,6-dimethoxy-4-methylphenyl)methanol 68: A solution of phenyl lithium was prepared by the addition of 11.85 g. (75.4 mmol.) of bromobenzene in 60 ml. of ether to 1.21 g. (174.1 mg. atom) of lithium and 30 ml. of ether. 10 g. (65.8 mmol.) of dimethoxy toluene was added and the reaction mixture was allowed to stand at room temperature under nitrogen for 60 hours. 1.95 g. (21.7 mmol.) of dimethyl carbonate in 200 ml. of benzene was added dropwise and the reaction mixture was refluxed for 3 days under nitrogen. The reaction mixture was poured into 200 ml. of water and the organic phase was concentrated to yield a gray residue. Recrystallization from ether gave 5.9 g. of the carbinol (56 %). 1H NMR (CD3CN, 299.95 MHz): 2.23 (s, 9H), 3.37 (s, 12H), 6.32 (s, 6H) 6.54 (s, 1H)13C NMR (75.43 MHz): 21.37, 56.67, 107.67, 118.30, 125.30, 136.78, 159.32. 219 Tris-g2,6-dimethoxy-4-phenylphenyl)methyl 69: 2-hydrox3_L—4-oxo-6-phenyl-2-cyclohexenecarboxylate : To a solution of 3.5 g. of NaOH in 150 ml. of methanol, 9.8 g. of dimethyl malonate (75 mmol.) was added and the mixture stirred for 30 min. at room temperature under argon. 10 g. (68 mmol.) of trans-4-phenyl-3~buten-2-one was added portion wise with stirring. The reaction mixture was refluxed for 3 hours under argon and allowed to cool to room temperature. The solvent was distilled under reduced pressure. The oily residue was treated with acidic (pH=5) water and extracted with dichloromethane. The solution was dried over magnesium sulfate, the solvent was evaporated, and the product mixture was dried under vacuum. The crude mixture was submitted to the next reaction without purification. 5-Phenyl Resorcinol : 20 g. (81.3 mmol.) of crude phenylcarboxylate 2~ hydroxy-4-oxo-6-pheny1-2-cyclohexenecarboxylate was dissolved in 150 ml. of DMF and 4.13 ml. (12.8g, 80 mmol.) of bromine was added dropwise with dry ice cooling under argon. The mixture was heated for three hours at about 80 °C until the evolution of C02 ceased followed by continued refluxing for 2 days. The solvent was distilled off under reduced pressure and the mixture was extracted with 300 ml. of 5% NaOH solution twice. The combined solution was treated with conc. HCl carefully with cooling and the 5-Phenyl resorcinol mixture was extracted with 500 ml. of methylene chloride. The solution was dried with magnesium sulfate and the solvent was evaporated under vacuum to give crystalline product. (5.6g, 43%) 1H NMR (CD3CN, 299.95 MHz): 6.36 (t, J=2 Hz, 1H), 6.61 220 (d, J=2 Hz, 2H), 7.31 (m, 1H), 7.40 (m, 2H), 7.55 (m, 2H), 8.35 (s, 2H), 13C NMR (CD3CN, 75.43 MHz): 102.48, 106.32, 107.11, 127.56, 128.10, 129.50, 144.05, 159.77. 3,5-dimethoxy-phenyl benzene : 10 g. (53.76 mmol.) of 5- phenylresorcinol was dissolved in 300 ml. of 5% NaOH solution, 12.72 ml. (16.95g, 134.4 mmol.) of dimethylsulfate was added using a separatory funnel, the mixture was mixed vigorously with shaking, and left for 30 min. followed by an extraction with 200 ml. of diethyl ether. The organic layer was dried, the solvent was evaporated, and the mixture was chromatographed over a silica gel using 30% ethyl acetate in hexane to give NMR clean product (8.9g, 78%).1H NMR (CDCl3, 299.95 MHz): 3.84 (s, 6H), 6.46 (t, J=2 Hz, 1H), 6.73 (d, J=2 Hz, 2H), 7.34 (m, 1H), 7.42 (tm, 2H), 7.56 (dm, 2H) 13C NMR (75.43 MHz): 54.99, 98.83, 105.01, 126.77, 127.13, 128.27, 140.76, 143.01, 160.59. Tris-(2,6-dimethoxy-4-phenyl phenyl)methanol : To a cooled solution of 2 g. (9.345 mmol.) of 5-phenyl—dimethoxybenzene in 100ml of diethyl ether, 4.11 ml. of 2.5 M n-BuLi in n-hexane was added slowly at -78°C under argon with stirring. The mixture was stirred at the room temperature for 2 days, 279 mg. (3.10 mmol.) of dimethyl carbonate in 100 ml. of benzene was added, and the reaction mixture was refluxed for 3 days under nitrogen. The reaction mixture was poured into 100 ml. of water, the organic layer was separated, dried with sodium sulfate, and the solvent was evaporated (4.4g, 71 %). 1H NMR (CDC13, 299.95 MHz): 3.54 (s, 18H), 6.75 (8, 6H), 6.95 (S, 1H), 7.32 (m, 3H), 7.41 (m, 6H), 7.59 (m, 6H) 13C 221 NMR (75.43 MHz): 56.43, 56.48, 105.02, 125.85, 126.83, 127.12, 128.64, 139.23, 141.40, 157.00, 158.79 m/z 668 (M+), found 668. A solution of carbinol in diethyl ether was treated with a small amount of conc. HBF4 and the blue precipitates were filtered and dried to give the cation tetrafluoroborate salt as crystalline solids. Cation: 1H NMR (CDC13, 299.95 MHz): 3.71 (s, 18H), 6.78 (s, 6H), 7.52 (m, 9H), 7.75 ((1, 1:7 Hz, 6H) 13C NMR (75.43 MHz): 56.48, 102.97, 126.95, 128.87, 129.94, 138.43, 192.29. A solution of the radical in diethyl ether was prepared according to general Method A. Di- A solution of phenyl lithium was prepared by the addition of 19.6 g. (125 mmol.) of bromobenzene in 60 ml. of ether to 2 g. (289 mol.) of lithium and 30 ml. of ether. 15g (109.6 mmol.) of 1,3-dimethoxybenzene was added and the reaction mixture was allowed to stand at room temperature under nitrogen for 60 hours. 20.3 g. (96.8 mmol.) of 2,6-dimethoxy,5 methyl benzoate in 200 ml. of benzene was added and the reaction mixture was refluxed for 3 days under argon. The reaction mixture was poured into 200 ml. of water and the organic phase was concentrated to yield gray residues. Recrystallization from ether gave a crude carbinol. A solution of the carbinol in diethyl ether was treated with a small amount of cone. HBF4 and the blue precipitate was filtered and dried to give the cation tetrafluoroborate salt as crystalline solids (12g, 27%). Cation: 1H NMR (CDC13, 299.95 MHZ): 2.48 (S, 3H), 3.54 (s, 12H), 3,59 (S, 6H), 6.36 (S, 222 2H), 6.48 (d, J=9 Hz, 4H), 7.41 (t, J=9 Hz, 2H) 13C NMR (75.43 MHz): 24.65, 56.60, 57.14, 104.75, 106.21, 138.03, 152.20, 160.63, 164.83, 165.71. Tris-12,6-dimethoxyphenyl) methyl ammonium tetrafluoroborate 72: To a 10 m1. THF solution of tris-(2,6-dimethoxyphenyl)methanol (500 mg., 1 mmol.), 0.2 ml. of cone. HBF4 was added followed by the addition of conc. aqueous ammonia until the color of the mixture turned red. 30 m1. of diethyl ether was added and the mixture was left to give red precipitates which were filtered and dried to give NMR (and X-ray) clean crystals of the product (380 mg., 72%). 1H NMR (CDC13, 299.95 MHz): 3.44 (s, 18H), 6.58 (d, J=8 Hz, 6H), 7.21 (t, J=8 Hz, 3H) 8.45 (s, 3H) 13C NMR (75.43 MHz): 56.25, 105.93, 106.02, 117.5, 129.06, 145.50, 157.50. Tris-(2,6-dimethoxyphenyl)methane 73 : To a solution of 5 g. (6.82 mmol.) of tris-(2,6-dimethoxyphenyl)methanol in 100 m1. of THF, 30 ml. of 10% HCl solution, 496.4 mg. (7.90 mmol.) of solid NaBH3CN was added slowly and the mixture was stirred for 2 hours at the room temperature. 100 ml. of water was added to the mixture and the product filtered, washed with water, and dried under vacuum to give 2.7 g. of clean products (93.4%). 1H NMR( CDCl3, 299.95 MHz): 3.41 (s, 18H), 6.45 (s, 1H), 6.48 (d, J=8 Hz, 6H), 7.02 (t, J=8 Hz, 3H) 13C NMR (75.43 MHz): 56.82, 106.37, 124.00, 125.64, 159.51 m/z 424 (M+), found 424. Tris-12,6-dimethoxyphenyl) silane 74: A solution of phenyl lithium was prepared by the addition of bromobenzene (39.2g., 0.250 mole) in 125 ml. of ether to lithium wire (4.0 g., 0.577 mol..) and 60 ml. of ether. 223 Resorcinol dimethyl ether (30.0 g., 0.218 mole) was added and the reaction mixture was allowed to stand at room temperature under nitrogen for 60 hours. 11.8 g. of triethoxysilane (0.072 mole) in 400 ml. of benzene was added and the reaction mixture was refluxed for 3 days under nitrogen. The reaction mixture was poured into 600 ml. of water and the organic phase was concentrated to yield gray residues. Recrystallization from diethyl ether gave 13.4 g. of the silane (42%). 1H NMR( CDC13, 299.95 MHz): 3.46 (s, 18H), 5.52 (s, 1H), 6.44 (d, J=8 Hz, 6H), 7.19 (t, J=8 Hz, 3H) 13C NMR (75.43 MHz): 55.41, 77.1, 103.80, 103.80, 129.69, 164.90. Tris-( 2,6-dimethoxyphenyl-3 ,5-dichlorophenyl )methyl ammonium tetra- fluoroborate 75: To a solution of 5 g. of tris-(2,6-dimethoxyphenyl)- methyl ammonium tetrafluoroborate in 10 ml. of diethyl ether and 50 ml. of dichloromethane, 10 ml. of sulfurylchloride was added and the mixture was refluxed under argon overnight. The organic layer was separated, dried with magnesium sulfate, the solvent was evaporated, and chromatographed over silica gel using methylene chloride to yield 2.75 g. of hexachloro-ammonium tetrafluoroborate as white crystals (37.5 %). 1H NMR (CDC13, 299.95 MHz): 3.38 (s, 9H), 3.66 (s. 9H), 7.53 (s, 3H), 8.69 (s, 3H). Tris-12,6-dimethoxyphenyl-3,5-dichlorophenyl) methane 76: To a solution of the methane 73 (2 g., 4.71 mmol.) in 10 ml. of diethyl ether and 50 ml. of dichloromethane, 33 mmol. of sulfurylchloride was added and the mixture was refluxed under argon overnight. The organic layer was separated, dried with magnesium sulfate, and the solvent was evaporated. Chromatography over silica gel using dichloromethane gave 224 hexachloromethane as white crystals (2.2g, 75%). 1H NMR (CDC13, 299.95 MHz): 3.07 (s, 9H), 3.73 (s, 9H), 6.65 (s, 1H), 7.31 (s, 3H) 13C NMR (75.43 MHz): 59.34, 61.00, 122.43, 123.94, 129.43, 133.32, 153.05, 156.93. Dim-dimethoxyphenyl)-_9Lhenyl methanol 77 : A solution of dimethoxy- phenyllithium was prepared by the addition of 34.8 ml. (86.9 mmol.) of 2.5 M n-BuLi in n-hexane to a solution of 10 g. (72.4 mmol.) of dimethoxybenzene in sodium-dried diethyl ether at -78°C and the mixture was stirred at room temperature under argon for 48 hours. 3.13 g. (23.0 mmol.) of methyl benzoate in 200 ml. of benzene was added and the reaction mixture was refluxed for 3 days under argon. The reaction mixture was poured into 200 m1. of water, the organic phase was washed with 200 ml. of water twice, dried, and concentrated to yield carbinol as gray residues. The carbinol was crystallized in dichloromethane and hexane to give 9 g. of NMR clean carbinol (65%). A solution of carbinol (500 mg., 1.31 mmol.), in 150 ml. of diethyl ether was treated with 0.1 m1. of cone. HBF4 and the blue precipitates were filtered and dried to give the cation tetrafluoroborate salt as a crystalline solid quantitatively. 200 mg. of the cation tetrafluoroborate salt in 20 ml. of dichloromethane was reduced by 20 ml. of 5M CrC12 solution (prepared with CrCl3-6H20 and Zn/Hg in 10% HCl solution) and extracted with dichloromethane and washed with water twice, dried with magnesium sulfate, and chromatographed with diethyl ether over silica gel to give a chromatographically pure solution of the radical (35%, based on SQUID measurement, neff). 1H NMR (CDC13, 299.95 MHz): 3.38 (s, 12H), 6.43 (s, 1H), 6.54 (d, J=8 Hz, 4H), 7.11 (t, J=8 Hz, 2H), 7.13 (m, 1H), 7.21 (m, 2H), 7.46 (m, 2H) 13C NMR 225 (75.43 MHz): 56.50, 78.50, 106.74, 125.37, 126.31, 126.50, 126.63, 127.28, 149.50, 158.00, mp. =104 -105°C. Cation: 1H NMR (CDC13, 299.95 MHz): 3.54 (s, 12H), 6.63 (d, J=8 Hz, 4H), 7.44 (m, 4H), 7.61 (m, 1H), 7.79 (t, J=8 Hz, 2H) 13C NMR (75.43 MHz): 56.83, 105.36, 125.00, 129.09, 134.59, 137.02, 145.90, 163.79. 191.58 m/e=364.44116 mp. =142 —144°C. Di-(2,6-dimethoxyphenyl)(4-chlorophenyl)-methanol 78: A solution of phenyl lithium was prepared by the addition of bromobenzene (39.2g., 0.250 mole) in 125 ml. of ether to lithium wire (4.0 g., 0.577 mol.) and 60 ml. of ether. Resorcinol dimethyl ether (30.0 g., 0.218 mole) was added and the reaction mixture was allowed to stand at room temperature under nitrogen for 60 hours. 14.8 g. (86.8 mmol.) of methyl 4-chlorobenzoate in 400 ml. of benzene was added and the reaction mixture was refluxed for 3 days under nitrogen. The reaction mixture was poured into 600 ml. of water and the organic phase was concentrated to yield gray residues. Recrystallization from methylene chloride in n-hexane gave 31.3 g. of NMR clean crystals of carbinol (87%). 1H NMR (CDC13, 299.95 MHz): 3.39 (s, 12H), 6.49 (s, 1H), 6.53 (d, J=8 Hz, 4H), 7.12 (t, J=8 Hz, 2H), 7.15 (dq, J=8 Hz, 2H), 7.37 (dq, J=8 Hz, 2H) 13C NMR (75.43 MHz): 56.29, 79.12, 106.625, 125.73, 126.64, 127.53, 128.11, 130.88, 148.01, 157.93. Cation 1H NMR (CDCl3, 299.95 MHz): 3.55 (s, 12H), 6.62 (d, J=8 Hz, 4H), 7.40 (dd, 4H), 7.78 (t, J=8 Hz, 2H) 13C NMR (75.43 MHz): 56.87, 105.42, 124.40, 129.50, 135.54, 142.65, 144.04, 146.23, 163.71, 189.04. 226 Di-(2,6-dimethoxy phenyl)(4-methoxyphenylzmethanol 79: A solution of phenyl lithium was prepared by the addition of bromobenzene (39.2 g., 0.250 mole) in 125 ml. of ether to lithium wire (4.0 g., 0.577 mol.) and 60 ml. of ether. Resorcinol dimethyl ether (30.0 g., 0.218 mole) was added and the reaction mixture was allowed to stand at room temperature under argon for 60 hours. 14.4 g. (98.6 mmol.) of methyl 4-methoxybenzoate in 400 ml. of benzene was added and the reaction mixture was refluxed for 3 days under argon. The reaction mixture was poured into 600 ml. of water and the organic phase was concentrated to yield gray residues. Recrystallization from methylene chloride in n—hexane gave 24.6 g. of NMR clean crystals of carbinol (61%). 1H NMR (CDC13, 299.95 MHz):3.38 (s, 12H), 3.76 (s, 3H), 6.36 (s, 1H), 6.53 (d, J=8 Hz, 4H), 6.75 (dq, J=8 Hz, 2H), 7.10 (t, J=8 Hz, 2H), 7.34 (dq, J=8 Hz, 2H) 13C NMR (75.43 MHz): 55.12, 56.47, 79.17, 106.88, 111.96, 126.65, 127.20, 127.73, 141.60, 157.40, 158.03, mp=105 0C. A solution of carbinol in diethyl ether was treated with a small amount of cone. HBF4 and the blue precipitates were filtered and dried to give the cation tetrafluoroborate salt as a crystalline solid. The cation tetrafluoroborate salt was reduced by CrClz solution (prepared with CrC13-6H20 and Zn/Hg in 10% HCl solution) and extracted with dichloromethane, washed with water twice, dried with magnesium sulfate, and chromatographed using diethyl ether over silica gel to give a chromatographically pure solution of radical (76%). The radical solution was left under the air for 3 days to give crystals of bis-(2,6-dimethoxy phenyl)methyl quinone methide shown above. 1H NMR (CDC13): 3.62 (s, 12H), 6.29 (dq, J=8 Hz, 2H), 6.56 (d, J=8 Hz, 4H), 7.20 (dq, J=8 Hz, 2H), 7.24 (t, J=8 Hz, 2H) 13C NMR (75.43 MHz): 55.98, 104.17, 117.5, 127.51, 130.24, 139.56, 150.00, 158.36, 188.00. Di- 2 6-dimethox hen l 4-meth 1 hen lmethanol : A solution of phenyl lithium was prepared by the addition of bromobenzene (39.2g., 0.250 mole) in 125 ml. of ether to lithium wire (4.0 g., 0.577 mol.) and 60 ml. of ether. Resorcinol dimethyl ether (30.0 g., 0.218 mole) was added and the reaction mixture was allowed to stand at room temperature under nitrogen for 60 hours. 14.8 g. (98.6 mmol.) of methyl 4-methylbenzoate in 400 m1. of benzene was added and the reaction mixture was refluxed for 3 days under nitrogen. The reaction mixture was poured into 600 ml. of water and the organic phase was concentrated to yield gray residues. Recrystallization from methylene chloride in n-hexane gave 32 g. of NMR clean crystals of carbinol (83%). 1H NMR (CDCl3, 299.95 MHz):2.29 (s, 3H), 3.38 (s, 12H), 6.38 (s, 1H), 6.53 (d, J=8 Hz, 4H), 7.00 (d, J=8 Hz, 2H), 7.10 (t, J=8 Hz, 2H), 7.32 (d, J=8 Hz, 2H) 13C NMR (75.43 MHz): 21.00, 56.46, 79.38, 106.88, 126.47, 126.61, 127.19, 127.37, 134.65, 146.23, 158.07. 228 Di-g 2,6-dimethoxy phenyl)(4-hydroxyphenyl)methanol 81: Bis-(2,6- dimethoxy phenyl)methyl quinone in THF was treated with conc. HBF4 and crystallized by slow evaporation to give needles of di-(2,6-dimethoxy phenyl)(4-hydroxyphenyl)methyl cation tetrafluoroborate (60% by nmr). 1H NMR (CDCl3): 3.56 (s, 12H), 6.56 (d, J=8 Hz, 4H), 7.17 (dq, =8 Hz, 2H), 7.46 (t, J=8 Hz, 2H), 7.82 (dq, J=8 Hz, 2H). Di-(2,6-dimethoxy phenyl)(4-nitrophenyl)methanol 83: A solution of dimethoxyphenyllithium was prepared by the addition of 70 ml. (175 mmol.) of 2.5 M n-Butyl lithium in n-hexane to a solution of 20 g. (145 mmol.) of dimethoxybenzene in sodium-dried diethyl ether at -780C and the mixture was stirred at room temperature under argon for 48 hours. 6.56 g. (36.2 mmol.) of methyl 4-nitrobenzoate in 300 ml. of benzene was added and the reaction mixture was refluxed for 3 days under nitrogen. The reaction mixture was poured into 300 m1. of water and the organic phase was dried and concentrated to yield brown residues. A solution of carbinol in diethyl ether was treated with a small amount of conc. HBF4 and the blue precipitates were filtered and dried to give the cation tetrafluoroborate salt as a crystalline solid. (<1%) Cation: 1H NMR (CDCl3, 299.95 MHz):3.56 (s, 12H), 6.63 (d, J=9 Hz, 4H), 7.40 (dd, J=8 Hz, 4H), 7.79 (t, J=9 Hz, 2H) 13C NMR (75.43 MHz): 56.88, 105.43, 124.41., 129.52, 135.55, 142.66, 144.01, 146.25, 163.73, 189.05. Di-(2,6-dimethoxy phenylH3,5-dimethoxyphenyl)methanol 84: A solution of phenyl lithium was prepared by the addition of bromobenzene (39.2g., 0.250 mole) in 125 ml. of ether to lithium wire (4.0 g., 0.577 mol..) and 60 ml. of ether. Resorcinol dimethyl ether (30.0 g., 0.218 229 mole) was added and the reaction mixture was allowed to stand at room temperature under nitrogen for 60 hours. 17 g. (98.6 mmol.) of methyl 3,5-dimethoxybenzoate in 400 ml. of benzene was added and the reaction mixture was refluxed for 3 days under nitrogen. The reaction mixture was poured into 600 ml. of water and the organic phase was concentrated to yield gray residues. Recrystallization from methylene chloride in n-hexane gave 24 g. of NMR clean crystals of carbinol. (55%). 1H NMR (CDCl3, 299.95 MHz): 3.40 (s, 12H), 3.70 (s, 6H), 6.27 (t, J=2 Hz, 1H), 6.46 (s, 1H), 6.52 (d, J=8 Hz, 4H), 6.69 (d, J=2 Hz, 2H), 7.09 (t, J=8 Hz, 2H), 13C NMR (75.43 MHz): 55.17, 56.43, 97.75, 105.25, 106.76, 127.28, 152.21, 158.06, 159.50. Tris-1 6-methoxy-2-g 2-methoxyethoxy )phenyl) methanol 85: 1-(2-methoxyethoxy)-3-methoxy benzene : A solution of 50 g. (400 mmol.) of methoxyphenol, 38.12 g. (400 mmol.) of chloromethoxyethane, 38.12g (400 mmol.) of chloromethoxyethane in 700ml of acetone was refluxed overnight under argon. The solution was filtered, evaporated, extracted with diethyl ether, dried with magnesium sulfate, concentrated, and chromatographed using 10% ethyl acetate in n-hexane to give the product as a clear oil (60%). 1H NMR (CDCl3): 3.42 (d, J=2 Hz, 3H), 3.70 (d, J=2 Hz, 2H), 3.74 (d, J=2 Hz, 3H), 4.07 (d, J=2 Hz, 2H), 6.50 (m, 3H), 7.16 (t, J=2 Hz, 1H) 13C NMR (75.43 MHz): 54.48, 58.45, 66.55, 70.37, 100.59, 106.01, 129.45, 159.75, 160.53. m/z 182 (M+), found 182. To a cooled solution of 14.82 g. (81.4 mmol.) of methoxyethoxy benzene in 100 ml. of diethyl ether, 39.1 ml. (97.7 mmol.) of 2.5 M n-BuLi in n- 230 hexane was added slowly at -78°C under argon with stirring. The mixture was stirred at room temperature for 2 days and 2.42 g. (26.86 mmol.) of dimethyl carbonate in 200 m1. of benzene was added and the reaction mixture was refluxed for 3 days under nitrogen. The reaction mixture was poured into 300 ml. of water and the organic layer was separated, dried with sodium sulfate, and the solvent was evaporated. A solution of carbinol in diethyl ether was treated with a small amount of conc. HBF4 and the blue precipitates were filtered and dried to give the cation tetrafluoroborate salt as a crystalline solid. The cation tetrafluoroborate salt was reduced by CrC12 solution (prepared with CrCl3-6H20 and Zn/Hg in 10% HCl solution), extracted with dichloromethane, washed with water twice, dried with magnesium sulfate, and chromatographed with diethyl ether over silica gel to give a chromatographically pure solution of the radical. 1H NMR (CDCl3): 3.17 (t, 6H), 3.19 (s, 9H), 3.39 (s, 9H), 3.81 (t, J=2 Hz, 6H), 6.47 (dd, 6H), 6.63 (s, 1H), 7.04 (t, J=2 Hz, 3H) 13C NMR (75.43 MHz): 55.68, 58.44, 68.27, 70.31, 105.93, 107.18, 125.78, 157.09, 158.57. Cation:1H NMR (CDC13): 3.14 (s, 9H), 3.21 (s, 6H), 3.55 (s, 9H), 3.93 (s, 6H), 6.50 (d, J=2 Hz, 6H), 7.57 (t, J=2 Hz, 3H) 13C NMR (75.43 MHz): 56.3, 58.4, 68.5, 69.7, 105.3, 121.0, 142.0, 156.6. Tris- 2 6-di 2-methox ethox hen l methanol 8 : l,3-di(2-methoxyethoxy2benzene : A solution of 70 g. (636 mmol.) of resorcinol, 185.2 g. (1340 mmol.) of potassium carbonate, 126.2 g. (1340 mmol.) of 2-chloroethyl methyl ether, 20 g. (120 mmol.) of potassium 231 iodide in 700 ml. of acetone was refluxed overnight under argon. The solution was filtered, evaporated, extracted with diethyl ether, dried with magnesium sulfate, concentrated, and chromatographed with 10% ethyl acetate in n-hexane to give the product as a clear oil (42.4%). 1H NMR (CDC13, 299.95 MHz) 3.42 (s, 6H), 3.72 (m, 4H), 4.07 (m, 4H), 6.51 (m, 3H), 7.13 (m, 1H) 13C NMR (75.43 MHz): 58.75, 66.74, 70.54, 101.24, 106.65, 129.34, 159.48. m/z 226 (M+), found 226. Tris-1 2,6-di( 2-methoxyethoxy )phenyl )methanol: 1,3 -di(2-methoxyethoxy)- benzene was prepared by the reaction of resorcinol and 2-chloroethyl methyl ether with K2CO3 in refluxing acetone. A solution of 1,3-di(2- methoxyethoxy) phenyl lithium was prepared by the addition of n-Butyl lithium in n-hexane to a solution of 1,3-di-(2-methoxyethoxy)benzene in sodium~dried diethyl ether at —78°C and the mixture was stirred at room temperature under argon for 48 hours. Dimethyl carbonate in benzene was added and the reaction mixture was refluxed for 3 days under argon. The reaction mixture was poured into ice water and the organic phase was concentrated to yield an an oil (11.9%). 1H NMR (CDC13, 299.95 MHz): 3.15 (s, 18H), 3.20 (m, 12H), 3.92 (m, 12H), 6.50 (d, J=9 Hz, 6H), 7.56 (t, J=9 Hz, 3H) 13C NMR (75.43 MHz): 58.37, 68.39, 69.80, 105.35, 111.50, 141.19, 161.82. A solution of the carbinol in diethyl ether was treated with a small amount of cone. HBF4 and the blue precipitates were filtered and dried to give the cation tetrafluoroborate salt as a crystalline solid. The cation tetrafluoro- borate salt was reduced by CrC12 solution (prepared with CrCl3-6HzO and 232 Zn/Hg in 10% HCl solution), extracted with dichloromethane, washed with water twice, dried with magnesium sulfate, and chromatographed with diethyl ether over silica gel to give a chromatographically pure solution of the radical 86. Bis-(2,6-dimethoxyphenyl)-4-chlorophenyl amine 87: The mixture of 950 mg. (7.6 mmol.) of 4-chloroaniline, 5 g. (18.9 mmol.) of 2,6—dimethoxy- iodobenzene, 6.9 g. (50 mmol.) of potassium carbonate, 2.54 g. (40 mg. atom) of copper, and 248 mg. (0.767 mmol.) of tris-(2-methoxyethoxy) amine was refluxed in 50 ml. of o-dichlorobenzene for 48 hours under argon. Inorganic salts and copper were removed by filtration of the hot reaction mixture. The solvent was distilled off under reduced pressure and the residue freed from the impurities by passing through a short silica gel column. Chromatography over silica gel using 20% ethyl acetate in hexane gave the product as white crystals (2.4g, 78%). 1H NMR (CDC13, 299.95 MHz): 3.58 (s, 12H), 6.40 (d, J=8 Hz, 2H), 6.56 (d, J=8 Hz, 4H), 6.97 (d, J=2 Hz, 2H), 7.20 (t, J=8 Hz, 2H) 13C NMR (75.43 MHz): 56.16, 104.00, 105.51, 111.5, 115.69, 125.75, 127.79, 157.22. Bis-12,6-dimethoxyphenyl)-3,5-dimethoxyphenyl amine 88: The mixture of 1.16 g. (7.6 mmol.) dimethoxyaniline, 5 g. (18.9 mmol.) 2,6- dimethoxy-iodobenzene, 6.9 g. (50 mmol.) of potassium carbonate, 2.54 g. (40 mmol.) of copper, and 248 mg. (0.767 mmol.) of tris-(2- methoxyethoxy) amine was refluxed in 50 ml. of o-dichlorobenzene for 48 hours under argon. Inorganic salts and copper were removed by filtration of the hot reaction mixture. The solvent was distilled off under reduced pressure and the residue freed from the impurities by passing through a 233 short silica gel column. Chromatography over silica gel using 20 % ethyl acetate in hexane gave the product as white crystals. (2.6lg, 81%). 1H NMR (CDC13, 299.95 MHZ): 3.59 (s, 12H), 3.60 (S, 6H), 5.80 (d, J=2 Hz, 2H), 5.87 (t, J=2 Hz, 1H), 6.53 (d, J=8 Hz, 4H), 7.15 (t, J=8 Hz, 2H) 13C NMR (75.43 MHZ): 55.00, 56.50, 90.82, 93.66, 105.50, 125.70, 157.37, 160.78. Tris-g3,5-dimethoxyphenyl)amine 89: The mixture of 460 mg. (3.80 mmol.) of 3,5—dimethoxyaniline, 2.51 g. (9.50 mmol.) of 2,6-dimethoxy- iodobenzene, 4.21 g. (30.54 mmol.) of potassium carbonate, 970 mg. (15.30 mmol.) of copper, and 248 mg. (0.767 mmol.) of tris-(2- methoxyethoxy) amine was refluxed in 20 ml. of o-dichlorobenzene for 24 hours under argon. Inorganic salts and copper were removed by filtration of the hot reaction mixture. The solvent was distilled off under reduced pressure and the residue freed from the impurities by passing through a short silica gel column. Chromatography over silica gel using 20 % ethyl acetate in hexane gave the product as white crystals (880 mg., 55%). 1H NMR (CDC13, 299.95 MHz): 3.68 (s, 18H), 6.14 (t, J=2 Hz, 3H), 6.24 (d, J=8 Hz, 6H) 13C NMR (75.43 MHz): 54.8, 54.9, 95.0, 102.6, 148.6, 160.7 m/z 425 (M+), found 425. Jacobson's (head-to-tail) peroxy dimer 91: A fresh solution of radical 77 was allowed to stand for 3 days under air to give the peroxy dimers. The peroxy dimer was crystallized by slow evaporation of a diethyl ether solution of the radical (65%, based on SQUID measurement, ueff) (the peroxydimer can be made almost quantitatively with an 02 saturated 234 solution of the radical in methylene chloride). 1H NMR (CDC13, 299.95 MHz): 3.40 (s, 12H), 3.58 (d, J=3 Hz, 12H), 4.82 (m, 1H), 5.54 (dd, J=12,2 Hz, 2H), 6.12 (dd, J=10,2 Hz, 2H), 6.44 (dd, 1:82 Hz, 8H), 6.55 (m, 2H), 7.03 (t, J=8 Hz, 4H), 7.05 (m, 1H), 7.11 (t, J=3 Hz, 2H), 7.68 (d, J=8 Hz, 2H). mp=108-111 °C. Martin's (tail~to-tail) peroxy dimer 90: A fresh solution of radical 1 was allowed to stand for 3days under air to give tail-to-tail peroxy dimer. (7% yield, based on SQUID measurement of the best radical, neff) (20%, based on average SQUID measurement of the radical, neffxwe can increase the yield of the peroxydimer significantly with 02 saturated solution of the radical in methylene chloride.) 1H NMR (CDC13, 299.95 MHz): 3.13 (s, 12H), 3.51 (d, J=3 Hz, 24H), 5.05 (d, J=2 Hz, 4H), 5.33 (t, J=2 Hz, 2H), 6.39 (dd, J=2,2 Hz, 8H), 7.00 (t, J=8 Hz, 4H). mp=>300 °C. Di-g2,6-dimethoxyphenyl)(2-hydroxyphenylzmethyl carbinol 92: A solution of phenyl lithium was prepared by the addition of bromobenzene (19.6g., 0.12 mole) in 75 ml. of ether to lithium wire (2.0 g., 0.288 mol..) and 30 ml. of ether. Resorcinol dimethyl ether (15.0 g., 0.109 mole) was added and the reaction mixture was allowed to stand at room temperature under nitrogen for 60 hours. 8.5 g. (56 mmol.) of methyl salicylate in 300 ml. of benzene was added and the reaction mixture was refluxed for 3 days under nitrogen. The reaction mixture was poured into 400 ml. of water and the organic phase was concentrated to yield a gray residue. Recrystallization from methylene chloride in n-hexane gave 10.1 g. of NMR clean crystals of carbinol (45%). 1H NMR (CDC13, 299.95 MHz): 3.40 (S, 12H), 6.58 (d, J=8 Hz, 4H), 6.64 (m, 2H), 6.86 (dd, J=8,2 Hz, 1H), 235 7.09 (m, 1H), 7.16 (t, J=8 HZ, 2H), 9.15 (s, 1H) 13C NMR (75.43 MHZ): 56.62, 81.37, 107.00, 115.74, 118.51, 123.89, 127.21, 127.93, 128.00, 131.73, 157.37, 158.25. 2,6-dimethoxyphenyl lithium was prepared by the addition of 34.8 ml. (86.9 mmol.) of 2.5 M n-BuLi in n-hexane to a solution of 10 g. (72.4 mmol.) of dimethoxy benzene in sodium-dried diethyl ether at ~780C, and the mixture was stirred at room temperature under argon for 48 hours. 3.10 g. (16.0 mmol.) of dimethylisophthalate in 200 ml of benzene was added and the reaction mixture was refluxed for 3 days under argon. The reaction mixture was poured into 200 ml. of ice water and the organic layer was separated, dried with sodium sulfate, and concentrated to yield a gray residue. Recrystallization from ether gave 2.23 g. of the dicarbinol as white crystals (21%). 1H NMR (CDC13, 299.95 MHz): 3.26 (s, 24H), 6.37 (s, 2H), 6.30 (d, J=8 Hz, 8H), 7.02 (t, J=8 Hz, 4H), 7.08 (m, 1H), 7.32 (dd, J=8,2 Hz, 2H), 7.44 (t, J=2 Hz, 1H) 13C NMR (75.43 MHz): 55.77, 79.42, 106.23, 123.56, 124.53, 126.37, 126.53, 146.88, 157.60. Bis-1,4-(tetra-(2,6-dimethoxyphenyl)methyl dicarbinol 94: A solution of 2,6-dimethoxyphenyl lithium was prepared by the addition of 34.8 ml. (86.9 mmol.) of 2.5 M n-BuLi in n-hexane to a solution of 10.0 g. (72.4 mmol.) of dimethoxy benzene in sodium-dried diethyl ether at -78°C and the mixture was stirred at room temperature under argon for 48 hours. 3.10 g. (16.0 mmol.) of dimethyl terephthalate in 200ml of benzene was added and the reaction mixture was refluxed for 3 days under argon. The reaction mixture was poured into 200 ml. of ice water and the organic 236 layer was separated, dried with sodium sulfate, and concentrated to yield a gray residue. Recrystallization from ether gave dicarbinol as white crystals. (4g, 37 %). 1H NMR (CDC13, 299.95 MHz): 3.33 (s, 24H), 6.18 (s, 2H), 6.48 (d, J=8 Hz, 8H), 7.06 (t, J=8 Hz, 4H), 7 .16 (m, 2H), 7.34 (m, 4H) 13C NMR (75.43 MHz): 56.22, 79.50, 106.73, 124.50, 125.02, 126.95, 128.00, 128.50, 158.03. Ox -4 4' dicarbinol : A solution of phenyl lithium was prepared by the addition of bromobenzene (39.2g., 0.250 mol.) in 125 ml. of ether to lithium wire (4.0 g., 0.577 mol.) and 60 ml. of ether. Resorcinol dimethyl ether (30.0 g., 0.218 mole) was added and the reaction mixture was allowed to stand at room temperature under nitrogen for 60 hours. 14.3 g. ( 50 mmol.) of 4-oxy-dimethylbenzoate in 400 ml. of benzene was added and the reaction mixture was refluxed for 3 days under nitrogen. The reaction mixture was poured into 600 ml. of water and the organic phase was concentrated to yield a gray residue. Recrystallization from methylene chloride in n-hexane gave 12.8 g. of NMR clean crystals of the dicarbinol (33%). 1H NMR (CDC13, 299.95 MHz): 3.40 (s, 24H), 6.38 (s, 2H), 6.53 (d, J=8 Hz, 8H) 6.83 (d, J=8 Hz, 4H), 7.10 (t, J=8 Hz, 4H), 7.34 (d, J=8 Hz, 4H) 13C NMR (75.43 MHz): 56.45, 79.24, 106.84, 116.96, 126.48, 127.32, 127.89, 143.96, 155.41, 158.09. 237 Tris-1,3,5-(bis-2,6-dimethoxyphenyl)hydroxymethyl benzene 26: A solution of phenyl lithium was prepared by the addition of bromobenzene (3.92g., 0.0250 mole) in 12.5 ml. of ether to lithium wire (0.4 g., 0.058 mol..) and 10 ml. of ether. Resorcinol dimethyl ether (3.0 g., 0.0218 mole) was added and the reaction mixture was allowed to stand at room temperature under nitrogen for 60 hours. 1.5 g. (5.9 mmol.) of 1,3,5- trimethylbenzoate in 400 ml. of benzene was added and the reaction mixture was refluxed for 3 days under argon. The reaction mixture was poured into 100 ml. of water and the organic phase was concentrated to yield 1.2 g. of gray residue (21%). 1H NMR (CDC13, 299.95 MHz): 3.20 (s, 36H), 6.35 (d, J=8 Hz, 12H), 6.95 (t, J=8 Hz, 6H), 7.33 (s, 3H), 7 .45 (s, 3H). 4 4'-bi hen llinked dicarbinol 97: l-(3,5-dimethoxyphenyl135-dimethoxybenzene : A solution of 2 g. (7.6 mmol.) of 3,5-dimethoxy-iodobenzene, Mg (7.6 mmol.), CuI (7.6 mmol.), dibromoethane (0.01 mmol.) in 30 ml. of THF was refluxed for 3 hours under argon. Iodo-2,6-dimethoxybenzene was added and the reaction mixture was refluxed further overnight. The solution was filtered, evaporated, extracted with methylene chloride, dried with magnesium sulfate, concentrated, and chromatographed with 10% ethyl acetate in n- hexane to give 870 mg. of white solid (83%). 1H NMR (CDC13, 299.95 MHz): 3.82 (s, 12H), 6.45 (t, J=8 Hz, 2H), 6.69 (d, J=8 Hz, 4H) 13C NMR (75.43 MHz): 54.99, 99.02, 105.06, 143.00, 160.51 m/z 274 (M+), found 274. 238 1-( 3 ,5-dimethoxyphenyl )-3 ,5-dimethoxy dimethylbenzoate: 1-(3,5-di- methoxypheny1)-3,5—dimethoxybenzene was lithiated with n-BuLi in dry THF and quenched with 10 fold excess of dimethyl carbonate and washed with water twice and extracted with diethyl ether followed by drying and evaporation of solvent. Chromatography over silica gel using 20 % ethyl acetate in hexane gave the product as white solids (23%). 1H NMR (CDC13, 299.95 MHz): 3.87 (s, 12H), 3.91 (s, 6H), 6.64 (s, 4H) 13C NMR (75.43 MHz): 52.51, 56.21, 103.45, 144.76, 157.49, 166.72. 1,3-benzenedicarboxaldehyde : A solution of 1,3-phenylene diamine (50g, 379 mmol.), hexamethylene tetraamine (188.3g, 1.345 mol.) in 90 ml. of conc. HCl and 600 m1. of 50% acetic acid in a 2 liter round bottom flask was refluxed for 3 hours with stirring. The hot amber reaction mixture is poured into a flask and a solution of 56.2 g. of NaOH in 750 ml. of water was slowly added with stirring. The mixture was covered and allowed to stand overnight at 5°C. The product was collected and washed with 10 ml. of cold water and dried (37.2g, 73 %). 1H NMR (CDC13, 299.95 MHZ): 7.72 (t, J=3 Hz, 1H), 8.14 (dd, J=4,1 Hz, 2H), 8.36 (m, 1H), 10.12 (s, 2H) 13C NMR (75.43 MHz): 129.50, 130.60, 134.22, 136.54, 190.65. 4- 3- 3-oxo-1-buten l hen l -3-buten-2-one: To a stirred solution of 10 g. (57.4 mmol.) of 1,3-benzenedicarboxaldehyde in 200 m1. of water and 200 m1. of diethyl ether, 100 ml. of acetone and 5 g. of NaOH were added and the mixture was stirred at the room temperature for 2 days. The organic layer was separated by adding 200 ml. of diethyl ether and 100 ml. of acetone and dried with magnesium sulfate. The combined organic 239 solution was evaporated and chromatographed over silica gel using 20% ethyl acetate in n—hexane to give the diketone as white crystals. (5.5 g., 45%). 1H NMR (CDC13, 299.95 MHz): 2.37 (s, 6H), 6.73 (d, J=l6 Hz, 2H), 7.46 (m, 2H), 7.54 (m, 3H), 7.67 (s, 1H) 13C NMR (75.43 MHz): 27.70, 127.86, 127.93, 129.58, 129.80, 135.19, 142.22, 198.05 m/z 214 (M+), found 214. 2-hydroxy-6-(3-13—hydroxy-2-methoxycarbonyl-5-oxo-3-cyclohexenyl) phenyl)-4-oxo-2-cyclohexenecarboxylate: To a solution of sodium methoxide (2.3g, 100 mg. atom) in 200 ml. of methanol, dimethyl malonate (12.36g, 93.63 mmol.) was added and the mixture stirred for 30 min at room temperature under argon. 4-(3-(3-0xo-1—butenyl)phenyl)-3- buten—2—one (6.68g, 31.21 mmol.) was added portion wise with stirring. The reaction mixture was refluxed for 3 hours under argon and allowed to cool to room temperature. The solvent was distilled under reduced pressure, the oily residue was treated with acidic (pH=5) water, and extracted with methylene chloride. The solution was dried over magnesium sulfate, solvent was evaporated, and the product mixture was dried under vacuum. The crude mixture was crystallized with methanol (10.7g, 83%) m/z 414 (M+), found 414. 1,3-Bis-( 3 ,5-dimethoxyphenyl )benzene: 1H NMR (CDC13, 299.95 MHz): 3.84 (s, 12H), 6.47 (t, J=2 Hz, 2H), 6.75 (d, J=2 HZ, 4H), 7.50 (m, 3H), 7.74 (s, 1H). Di- 2 6-dimethox hen l - 4- rid l methanol 99: A solution of phenyl lithium was prepared by the addition of bromobenzene (39.2g., 0.250 240 mole) in 125 ml. of ether to lithium wire (4.0 g., 0.577 mol.) and 60 ml. of ether. Resorcinol dimethyl ether (30.0 g., 0.218 mole) was added and the reaction mixture was allowed to stand at room temperature under nitrogen for 60 hours. 13.5 g. ( 98.6 mmol.) of methyl isonicotinate in 400 m1. of benzene was added and the reaction mixture was refluxed for 3 days under nitrogen. The reaction mixture was poured into 600 ml. of water and the organic phase was concentrated to yield a gray residue. Recrystallization from methylene chloride in n-hexane gave 14 g. of NMR clean crystals of carbinol (37 %). 1H NMR (CDC13, 299.95 MHz): 3.39 (s, 12H), 6.52 (d, J=8 Hz, 4H), 6.65 (s, 1H), 7.13 (t, J=8 Hz, 2H), 7.36 (dd, J=8,2 Hz, 2H), 8.40 (dd, J=8,2 Hz, 2H) 13C NMR (75.43 MHz): 56.02, 78.80, 106.30, 121.55, 124.41, 127.82, 148.31, 157.74, 158.50. Di-(2,6-dimethoxy phenylz-(3-pyridyl) methanol 100: A solution of phenyl lithium was prepared by the addition of bromobenzene (39.2g., 0.250 mole) in 125 ml. of ether to lithium wire (4.0 g., 0.577 mol.) and 60 ml. of ether. Resorcinol dimethyl ether (30.0 g., 0.218 mole) was added and the reaction mixture was allowed to stand at room temperature under nitrogen for 60 hr. 13.5g ( 98.6 mmol.) of methyl nicotinate in 400 m1. of benzene was added and the reaction mixture was refluxed for 3 days under nitrogen. The reaction mixture was poured into 600 ml. of water and the organic phase was concentrated to yield a gray residue. Recrystallization from methylene chloride in n—hexane gave 15.5 g. of NMR clean crystals of the carbinol (41 %). 1H NMR (CDC13, 299.95 MHz): 3.40 (s, 12H), 6.51 (d, J=8 Hz, 4H), 6.63 (s, 1H), 7.12 (t, J=8 Hz, 3H), 7.70 (dt, J=8,2 Hz, 1H), 8.34 (dd, J=6,2 Hz, 1H), 8.64 (d, J=2 Hz, 1H) 13C NMR (75.43 MHz): 241 13C NMR (75.43 MHZ): 56.16, 78.20, 106.36, 121.73, 124.82, 127.76, 133.73, 144.46, 146.39, 148.93, 157.83. Di-(2,6-dimethoxy phenyl)(2-pyridyl) methanol 101: A solution of phenyl lithium was prepared by the addition of bromobenzene (39.2g., 0.250 mole) in 125 ml. of ether to lithium wire (4.0 g., 0.577 mol..) and 60 ml. of ether. Resorcinol dimethyl ether (30.0 g., 0.218 mole) was added and the reaction mixture was allowed to stand at room temperature under nitrogen for 60 hours. 13.5 g. ( 98.6 mmol.) of methyl picolinate in 400 ml. of benzene was added and the reaction mixture was refluxed for 3 days under nitrogen. The reaction mixture was poured into 600 ml. of water and the organic phase was concentrated to yield a gray residue. Recrystallization from methylene chloride in n-hexane gave 23 g. of NMR clean crystals of the carbinol (61 %). 1H NMR (CDC13, 299.95 MHz): 3.39 (s, 12H), 6.49 (d, J=8 Hz, 4H), 6.56 (s, 1H), 7.00(m, 1H), 7.09 (t, J=8 Hz, 2H), 7.50 (m, 2H), 8.45 (m, 1H) 13C NMR (75.43 MHz): 56.30, 79.44, 106.24, 120.45, 121.70, 127.32, 134.73, 146.71, 158.07, 168.46. 242 REFERENCES 1. C. J. Pedersen, J. Am. Chem. Soc. 89, 7017, 1967.D. J. Cram, Angew. Chem. Int. Ed. Engl., 25, 1039, 1986. D. J. Cram, Science, 240, 760, 1988. J.-M. Lehn, Angew. Chem. Int. Ed. Engl., 27, 89, 1988. 2 H. G. Viehe, R. Merényi, L. Stella, Z. Janousek, Angew. Chem. Int. Ed. Engl. 18, 917, 1979; M. Ballester, Acc. Chem. Res. 18, 380, 1985. 3 For carbenes with particular relevance to organic ferromagnets, see: A. M. Trozzolo, R. W. Murray, G. Smolinsky, W. A. Yager, E. Wasserman, J. Am. Chem. Soc. 85, 2526, 1963; K. Itoh, Chem. Phys. Lett. 1, 235, 1967; T. Takui, K. Itoh, Chem. Phys. Lett. 19, 120, 1973. A. Izuoka, S. Murata, T. Sugawara, H. Iwamura, J. Am. Chem. Soc. 107, 1786, 1985; Y. Teki, T. Takui, K. Itoh, H. Iwamura, K. Kobayashi, J. Am. Chem. Soc. 108, 2147, 1986; I. Fujita, Y. Teki, T. Takui, T. Kinoshita, K. Itoh, F. Miko, Y. Sawaki, H. Iwamura, A. Izuoka, T. Sugawara, J. Am. Chem. Soc. 112, 4074, 1990. 4 W. T. Borden, Ed.,Diradicals; John Wiley &' Sons, Inc., 1982; G. L. Closs, R. J. Miller, 0. D. Redwine, Acc. Chem. Res. 18, 196, 1985.; M. Platz (Ed.) Symposia-in~Print No. 4, Tetrahedron, 38, 733-867; J. A. Berson, Acc. Chem. Res. 11, 446, 1978.; C. Doubleday, Jr.; N. J. Turro, J.~F. Wang, Acc. Chem. Res. 22, 199, 1989. 5 For an excellent discussion of exchange phenomena in bimetallic systems see: R. D. Willett, D. Gatteschi, O, Kahn Ed.;"Magneto—Structural Correlations in Exchange-Coupled Systems," NATO ASI Series C, V. 140, D. Reidel, Boston, 1985.; O. Kahn, Structure and Bonding, 68, 91, 1987.; A. Bencini, D. Gatteschi, Electron Paramagnetic Resonance of Exchange Coupled Systems, Springer-Verlag, Berlin, 1990. 6 V. Bonacic'—Koutecky', J. Koutecky’, J. Michl, Angew. Chem. Int. Ed. Eng]. 26, 170, 1987. 7 H. McConnell, J. Chem. Phys. 39, 1910, 1963.; Proc. Robert A Welch Found. 108,11,1967. 8 C. M. Hurd Contemporary Physics 23, 469 1982. 9 197th National American Chemical Society meeting in Dallas, Texas, April 10-14, 1989. Proceedings of the Symposium on Ferromagnetic and High Spin Molecular Based Materials in Mol. Cryst. Liq. Cryst. 176, 1-562 (1989); J. S. Miller; D. A. Dougherty, Angew. Chem. Int. Ed. Engl., Adv. Mater. 28, 961 (1989). 10 w. E. Broderick, J. A. Thompson, E. P. Day, B. M. Hoffman, Science, 249, 402 (1990); J. S. Miller, A. J. Epstein, in: Z. Yoshida, T. Shiba, Y. Oshiro (Eds) New Aspects of Organic Chemistry I, VCH, Weinheim, 1989; p 237; J. S. ¥ 243 11 12 l3 14 15 16 17 18 19 Miller, A. J. Epstein, W. M. Reiff, Chem. Rev. 88, 201 (1988). J. S. Miller, A. J. Epstein, W. M. Reiff, Science 240, 40 (1988). Chem. Eng. News, p. 5 (Mar. 30, 1987) and p. 28 (Nov. 9, 1987). J. S. Miller, J. C. Calabrese, H. Rommelmann, S. R. Chittipeddi, J. J. Zhang, W. M. Reiff,'A. J. Epstein, J. Am. Chem. Soc. 109, 769 (1987). J. S. Miller, A. J. Epstein, J. Am. Chem. Soc. 109, 3850 (1987). P. W. Selwood, Magnetochemistry, Interscience, New York, 1943; A. P. Cracknell, Magnetism in Crystalline Materials, Pergamon Press, Oxford, 1975; R. L. Carlin, Magnetochemistry, Springer-Verlag, Berlin, 1986. For early musings on organic ferromagnetism see ref. 9 and: N. Mataga, Theoret. Chim. Acta 10, 372 (1968); A. A. Ovchinnikov, Theoret. Chim. Acta, 47, 297 (1978); A. L. Buchachenko, Doklady Nauk Acad. SSSR 244, 1146 (1979); I. A. Misurkin, A. A. Ovchinnikov, Russ. Chem. Rev. (Engl. Trans.) 46, 967 (1977); R. Breslow, B. Jaun, R. Q. Klutz, C.-Z. Xia, Tetrahedron, 38, 863 (1982). For some more recent magnetic models see: H. Fukutome; A. Takahashi; M. Ozaki, 133, 34 (1987); L. Dulog, J. S. Kim, Angew. Chem. Int. Ed. Engl., 29, 415 (1990).; K. Yamaguchi, Y. Toyoda, T. Fueno, Synthetic Metals, 19, 81 (1987).; K. Yamaguchi, Y. Toyoda, M. Nakano, T. Fueno, ibid., 19, 87 (1987).; K. Awaga, T. Sugano, M. Kinoshita, J. Chem. Phys., 85, 2211 (1986). K. Awaga, T. Sugano, M. Kinoshita, Chem. Phys. Lett., 141, 540 (1987). P. M. Lahti, A. S. lchimura, Mol. Cryst. Liq. Cryst., 176, 125 (1989). P. M. Lahti, A. S. lchimura, J. A. Berson, J. Org. Chem., 54, 958 (1989).; K. M. Chi, J. C. Calabrese, J. S. Miller, Mol. Cryst. Liq. Cryst., 176, 173 (1989). K. M. Chi, J. C. Calabrese, J. S. Miller, S. I. Khan, Mol.Cryst. Liq. Cryst., 176, 185 (1989). DJ. Cram, Angew. Chem. Int. Ed. Engl., 27, 1009 (1988).; J.-M. Lehn, Angew Chem. Int. Ed. Engl. 27, 89 (1988). CPK models are particularly faithful for ion binding ligands and their interactions and have consistently provided insights into complexant design. In the present case, two CPK models of radical l neatly encapsulate a marble of appropriate size to represent a Li+ ion. The methoxy groups in the analogous tris(2,6~dimethoxyphenyl)phosphine participate in the ligation of a variety of metal species. See: K.R. Dunbar, S. C. Haefner, L. E. Pence, J. Am. Chem. Soc., 111, 5504 (1989). J.-M. Lehn, Angew. Chem. Int. Ed. Engl., 27, 89, 1988. E. Weber, Top. Curr. Chem, 140, 1, 1987. S. Kamitori, K. Hirotsu and T. Higuchi, J. Am. Chem. Soc. , 109, 2409, 1987. W. Saenger, Angew. Chem, 92, 343, 1980, F. Vogtle and W. M. Miiller, Angew. Chem. Int. Ed. Engl., 18, 628, 1979. K. Odashima and K. Koga, in Cyclophanes, Vol. 2. Academic Press, New York, 629, 1983. K. Odashima, A. Ita, Y. Litaka and K. Koga, J. Org. Chem, 50, 4478, 1985. 244 20 21 22 23 24 25 26 27 28 29 30 31 32 33 F. Vogtle and W. M. Muller, Angew. Chem. 96, 711, 1984, F. V'dgtle and W. M. Miiller Angew. Chem. Int. Ed. Engl., 25, 567, 1986. H. J. Schneider and T. Blatter Angew. Chem. 100, 1211, 1988. H. J. Schneider and T. Blatter Angew, Chem. Int. Ed. Engl., 27, 1163, 1988. H.—J. Schneider and I. Theis, Angew. Chem, Int. Ed. Engl., 28, 753, 1989. H.-J. Schneider, D. Guttes and U. Schneider, J. Am. Chem. Soc. 110, 6449, 1988. F. Diederich and K. Dick, Angew. Chem. Int. Ed. Engl. 22, 715, 1983. F. Diederich and K. Dick, Angew. Chem. Int. Ed. Engl. 23, 810, 1984. F. Diederich and D. R. Carcanague, Angew. Chem. Int. Ed. Engl. 29, 769, 1990. G. R. Desiraju "Crystal Engineering: the Design of Organic Solids” Materials Science Monographs, 54, Elsevier, New York, NY. 1989. G. M. J. Schmidt, Pure Appl. C hem, 27 , 647, 1971. and references therein. H. R. Ott Nature 363, 56, 1993. M. A. Beno, H. H. Wang, K. D. Carson, A. M. Kini, G. M. Frankenbach, J. R. Ferraro, N. Larson, G. D. McCabe, J. Thomson, C. Purnama, M. Vashon, J. M. Williams, D. Jung, M. -H. Whangbo., Mol. Cryst. Liq. Cryst. 1990, 181, 145. M. A. Beno, H. H. Wang, A. M. Kini, K. D. Carson, U. Geiser, W. K. Kwok, J. E. Thompson, J. M. Williams, J. Ren, M. -H. Whangbo, Inorganic. Chem. 1990,29, 1599. D. Griller and K. U. Ingold Acc. Chem. Res. 9, 13, 1976. S. W. Benson "Thermochemical Kinetics", Wiley, New York, N.Y., 1968, D. M. Golden and S. W. Benson, Chem. Rev. 69, 125, 1969, H. E. O'Neal and S. W. Benson in ”Free Radicals”, Vol II. J. K. Kochi, Ed., Wiley, New York, NY. 197 3. "T he History of Organic Chemistry in the United States, J 875 -I 955 " D. S. Tarbell and A. T. Tarbell, K & S Press, Nashville, TN, 1986. H. Lankarnp, W. Th. Nauta, C. MacLean, Tett. Lett. 249, 1968. P. Jacobson, Ber. Dtsch. Chem. Ges. 38, 196, 1905. M. Stein, W. Winter, and A. Rieker Angew. Chem. Int. Ed. Engl. 17, 692, 1978. B. Kahr, D. V. Engen, and K. Mislow J. Am. Chem. Soc. 108, 8305, 1986. W. Schlenk and M. Brauns, Ber. Dtsch. Chem. Ges. 48, 661, 1915. G. Kothe, K.-H. Denkel, W. Sfimmermann Angew. Chem. Int. Engl. Ed.,9, 906, 1970. A. E. Chichibabin, Chem. Ber. 40, 1810, 1907. 245 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 M. S. Platz, Diradicals, W. T. Borden, Ed., Wiley, New York, 1982, Chapter 5. L. K. Montgomery, J. C. Huffman, E. A. Jurczak, M. P. Grendze, J. Am. Chem. Soc. 108, 6004, 1986. J. Thiele, H. Balhorn, Chem. Ber. 37, 1463, 1904. M. Ballester, C. Molinet, Chem. Ind. (London) 1290, 1954. M. Ballester, C. Molinet, J. Castaner, J. Am. Chem. Soc. , 82, 4254, 1960, M. Ballester, R. Riera, J. Am. Chem. Soc., 86, 4505, 1964. M. Ballester, Pure Appl. Chem. 15, 123, 1967. M. Ballester, J. Castaner, J. Riera, J. Am. Chem. Soc. 88, 957, 1966. M. J. Sabacky, C. S. Johnson, Jr., R. G. Smith, H. S. Gutowsky, J. C. Martin, J. Am. Chem. Soc. 89, 2054 (1967). While EPR spectra observed by Jackson and Kahr remain unassigned at least one feature is clear; the spectral dispersion—as large as 800 Gauss in some cases—is too great to be due solely to hyperfine interactions and must be fine structure resulting from long-range magnetic coupling of pairs of radicals. The temperature dependence of the signal intensity indicates that its carrier might be species with a high Spin ground state. The Spectra were unchanged when LiBF4 was replaced by LiClO4 or 99% 61.113134. For the status of research on molecular magnetism in 1985, see H. Iwamura, T. Sugawara, K. Itoh, T. Takui, Mol. Cryst. Liq. Cryst.125, 251 (1985) R. Breslow, Mol. Cryst. Liq. Cryst. 125, 261 (1985) For a valuable enumeration of these models, see D. Dougherty and J. S. Miller in Mol. Cryst. Liq. Cryst. 176, 25 (1989). S. Goldschmidt, Ber. 53, 44, 1920, S. Goldschmidt and K. Euler Ber. 55, 616, 1922. S. Goldschmidt and K. Renn, Ber. 55, 628, 1922. S. Goldschmidt, and F. Graef, Ber. 61, 1858, 1928. R. H. Poirier, E. J. Kahler, and F. Benington, J. Org. Chem. 17, 1437, 1952. L. S. Singer, and C. Kikuchi, J. Chem. Phys. 23, 1738, 1955. G. Fraenkel and PD. Bartlett, J. Am. Chem. Soc. 81, 5582, 1959. F. L. Allen and S. Sugden, J. Chem. Soc. 440, 1936. W Duffy, J. Chem. Phys. 36, 490, 1962, W. Duffy and D. L. Strandburg, J. Chem. Phys. 46. 457, 1967. C. A. Hutchison, R.C. Pastor, and AG. Kowalsky J. Chem. Phys. 20, 534, 1952. N. W. Lord and S. M. linder, J. Chem. Phys. 34, 1693, 1961. R. M. Deal and W. S. Koski, J. Chem. Phys. 31, 1138, 1959. R. W. Holmberg, R. Livingston, and W. T. Smith, J. Chem. Phys. 33, 541, 1960. R. 1. Walter, J. 246 49 50 51 52 53 54 55 56 57 58 59 60 61 Am. Chem. Soc. 88, 1923, 1966, R. 1. Walter, J. Am. Chem. Soc. 88, 1930, 1966. J. Heidberg and J. Weil, J. Am. Chem. Soc.86, 5173, 1964. A. L. Buchachenko, "Stable Radicals", Consultants Bureau, New York. 1965. O. Piloty and BC. Schwerin. Ber. 34, 1870, 1901. O. Piloty and B. G. Schwerin Ber. 34, 2354, 1901. K. H. Meyer and H. G. Bilbroth Ber. 52, 1476, 1919. K. H. Meyer and W. Reppe, Ber. 52, 1476, 1919. Y. Y. Lim and R. S. Drago, J. Am. Chem. Soc. 93, 891, 1970. Y. Y. Lim and R. S. Drago, Inorganic Chem. 11, 1334, 1972. R. Briere, H. Lemaire, and A. Rassat, Bull Soc. Chim. France, 3273, 1965. L. Chapelet, H. Lemaire, and A. Rassat, Bull. Soc. Chim. F ranmce, 3283, 1965. E. G. Rozantsev, Bull. Acad. Sci. USSR. 2085, 1964. A.M. Feldman and A. K. Hoffmann Chem. Abst. 61, 13289, 1964. A. W. Hanson, Acta Cryst. 6, 32, 1953. K. H. Meyer and W. Reppe Ber. 54, 327, 1921. A. K. Hoffmann, A. M. Feldman, and E. Gelblum, J. Am. Chem. Soc. 86, 646, 1964. C. D. Cook, J. Org. Chem. 18, 261, 1953. C. D. Cook and R. C. Woodworth, J. Am. Chem. Soc. 75, 6242, 1953. N. C. Yang and A. Castro, J. Am. Chem. Soc. 82, 6208, 1960. E. A. Chandross J. Am. Chem. Soc. 86, 1263, 1964. F. A. Neugebauer, H. Tischmann, and M. Jenne, Angew. Chem. Int. Ed., 6, 362, 1967. E. A. Chandross, J. Am. Chem. Soc. 86, 1263, 1964. K. Mukai and J. Sakamoto J. Chem. Phys, 68, 1432, 1978. A. Earnshaw, Introduction to Magnetochemistry, Academic Press, New York, 1968. R. L. Carlin , Magnetochemistty, Springer—Verlag, New York, 1986. B. Morosin and E. J. Graeber. Acta Cryst. 16, 1176, 1963. L. J. de Jongh and A. R. Miedema, Adv. Phys. 23, 1, 1974. D.W. Hone and P. M. Richards, Ann. Rev. Material Sci. 4, 337, 1974. R. Dingle, M. E. Lines, and S. L. Holt, Phys. Rev. 187, 643, 1969. T. Oguchi, Prog. Theor. Phys., 13, 148, 1955. H. Ohya-Nishiguchi, Bull. Chem. Soc. Japan. 52, 3480, 1979. J. C. Bonner, M. E. Fisher, Phys. Rev. 135, A640, 1964. J. C. Bonner, H. W. J. Blote, J. W. Bray, and I. S. Jacobs, J. App. Phys. 50, 1810, 1979. M. Verdaguer, A. Gleizes, J. P. Renard, and J. Seiden, Phys. Rev. B. 29, 5144, 1984. W. Duffy, Jr. and K. P. Barr, Phys. Rev. 165, 647, 1968. 247 62 63 64 65 66 67 68 69 70 71 72 73 74 75 M. E. Fisher, J. Phys. 32, 343, 1964. A. van der Bilt, K. O. Joung, and R. C. Carlin, Phys. Rev. B. 24, 445, 1981. M. M. Olmstead, W. K. Musker, L. W. Ter Haar, and W. E. Hatfield. J. Am. Chem. Soc. 104, 6627, 1982. J. W. Hall, W. E. Marsh, R. R. Weller, and W. E. Hatfield, Inorg. Chem. 20, 1033, 1981., W. E. Hatfield, R. R. Weller, and J. W. Hall, Inorg. Chem. 19, 3825, 1980., W. H. Crawford and W. E. Hatfield, Inorg. Chem. 16, 1336, 1977. A. Bencini, C. Benelli, D. Gatteschi, and C. Zanchini, J. Am. Chem. Soc. 106, 5813, 1984. C. Benelli, D. Gatteschi, D. W. Carnegie, Jr., and R. L. Carlin, J. Am. Chem. Soc. 107, 2560, 1985. A. S. Edelstein, M. Mandel, J. Chem. Phys. 35, 1130, 1961. H. M. McConnell J. Chem. Phys. 37, 1910, 1963. H. McConnell, Proc. of the Robert A. Welch Foundation C onf. on Chem. Res. 11, 144, 1967 . N. Mataga, Theoret. Chim. Acta (Berl.) 10, 372, 1968. A. A. Ovchinnikov, Theoret. Chim. Acta. (Berl.) 47, 297, 1978. A. A. Ovchinnikov, V. N. Spector Synthetic Metals, 27 , B615, 1988. R. Breslow Pure & Appl. Chem. 54, 927, 1982. R. Breslow, B. Jaun, R. Q. Kluttz, C.—Z. Xia Tetrahedron, 38, 863, 1982. T. LePage, R. Breslow, J. Am. Chem. Soc. 109, 6412, 1987. R. Breslow. Mol. Cryst. Liq. Cryst., 176, 199, 1989. J. B. Torrance, S. Oostra, A. Nazzal, Synthetic Metals, 19, 709, 1987. C. Kollmar, O. Kahn, J. Am. Chem. Soc. 113, 7987, 1991, C. Kollmar, M. Couty, and O. Kahn, J. Am. Chem. Soc. 113, 7994, 1991. C. Kollmar and O. Kahn, Acc. Chem. Res. 26, 259, 1993. A. Caneschi, D. Gatteschi, and R. Sessoli, Acc. Chem. Res. 22, 392, 1989, J. S. Miller, A. J. Epstein, and W. M. Reiff, Acc. Chem. Res. 21, 114, 1988, J. S. Miller, A. J. Epstein, and W. M. Reiff, Chem. Rev. 88, 201, 1988. O. Kahn, Angew. Chem. Int. Ed. Engl. 24, 834, 1985. B. Bleaney, K. D. Bowers, Proc. R. Soc. London A 214, 451, 1952. D. Gatteschi, O. Kahn, R. D. Willett (eds): Magneto Structural Correlations in Exchange Coupled Systems, D. Reidel, Dordrecht 1984. M. Verdaguer, M. Julve, A. Michalowicz, and O. Kahn, Inorg. Chem. 22, 2624, 1983. Y. Pei, O. Kahn, J. Sletten., J. Am. Chem. Soc. 108, 3143, 1986., Y. Pei, M. Verdaguer, O. Kahn, J. Sletten, and J. P Renard, J. Am. Chem. Soc. 108, 7428, 1986. Y. Pei, M. Verdaguer, O. Kahn, J. Sletten, and J. P. Renard, Inorg. Chem. 26, 138, 1987. O. Kahn, Angew. Chem. Int. Ed. Engl. 24, 834, 1985. K. Nakatani, O. Kahn. C. Mathoniere, T. Pei, and C. Zakine, New. J. 248 76 77 78 79 80 81 82 83 84 Chem. 14, 861, 1990. O. Kahn, Comments Inorg. Chem. 3, 105, 1984. M. Julve, M. Verdaguer, A. Gleizes, M. P. Levisalles, adn O. Kahn. Inorg. Chem. 23, 3808, 1984. D. Beltram, E. Escriva, and M. Drillon, J. Chem. Soc., Faraday Trans. 2, 78, 1773, 1982. M. Drillon, E. Coronado, D. Beltran, J. Curely, R. Georges, P. R. Nugteren, L. J. de Hongh, and J. L. Genicon, J. Magn. Magn. Mater, 54-57, 1507, 1986. M. Drilon, E. Coronado, A. Fuertes, D. Beltran, and R. Georges, Chem. Phys. 79, 449, 1983. E. Coronado, M. Drillon, A. Fuertes, D. Beltran, A. Mosset, and J. Galy, J. Am. Chem. Soc. 108, 900, 1986. O. Kahn Comments Inorg. Chem. 3, 105, 1984. R. D. Willett, D. Gatteschi, O. Kahn Eds., Magnetostructural Correlations in Exchange Coupled Systems (NATO Advanced Study Institute Series), D. Reidel, Dordrecht, The Netherlands, 1984. D. Y. Jeter, D. L. Lewis, J. C. Hemplel, D. J. Hodgson, and W. E. Hatfield, Inorga. Chem. 11, 1958, 1972. M. Julve, M. Verdaguer, A. Gleizes, M. P. Levisalle, and O. Kahn, Inorg. Chem. 23, 3808, 1984. Y. Pei, K. Nakatani, J. Sletten, J. P. Renard, Inorg. Chem. 16, 3170, 1986. L. R. Carlin, A. Kopinga, O. Kahn, and M. Verdaguer, Inorg. Chem. 25, 1786, 1986. S. K. Shakhatreh, E. G. Bakalbassis, I. Brfidgam, H. Hartl, J. Mrozinski, and C. A. Tsipis, Inorg. Chem. 30, 2801, 1991. A. Bencini, C. Benelli, D. Gatteschi, and C. Zanchini, J. Am. Chem. Soc. 106, 5813, 1984. C. Benelli, D. Gatteschi, and C. Zanchini, Inorg. Chem. 23, 798, 1984. C. Benelli, D. Gatteschi, D. W. Carnegie, Jr., and R. Carlin, J. Am. Chem. Soc. 107, 2560, 1985. A. Caneschi, D. Gatteschi, J. Laugier, P. Rey, R. Sessoli, and C. Zanchini, J. Am. Chem. Soc. 110, 2795, 1988. A. Caneschi, D. Gatteschi, J. Laugier, P. Rey, and C. Zanchini, Inorg. Chem. 28, 1969, 1989. A. Caneschi, D. Gatteschi, J. P. Renard, P. Rey, and R. Sessoli, Inorg. Chem. 28, 1976, 1989. A. Caneschi, D. Gatteschi, J. P. Renard, P. Rey, and R. Sessoli, Inorg. Chem. 28, 3314, 1989. A. Caneschi, D. Gatteschi, and R. Sessoli, Mol. Cryst. Liq. Cryst. 176, 329, 1989. A. Caneschi, D. Gatteschi, J. P. Renard, P. Rey, and R. Sessoli, J. Am. Chem. Soc. 111, 785, 1989. P. Rey, J. Laugier, A. Caneschi, D. Gatteschi, Mol. Cryst. Liq. Cryst. 176, 337, 1989. A. Caneschi, F. Ferraro, D. Gatteschi, P. Rey, and R. Sessoli, Inorg. Chem. 29, 4217, 1990. A. Caneschi, D. Gatteschi, R. Sessoli, and P. Rey, Acc. Chem. Res. 22, 392, 1989. A. Caneschi, D. Gatteschi, P. Rey, and R. Sessoli, Chem. Mater, 4, 204, 1992. S. S. Eaton, G. R. Eaton, Coord, Chem, Rev. 26, 207, 1978. S. Eaton, S. S. G. R. Eaton, Coord, Chem. Rev. 83, 29, 1988. A. Caneschi, D. Gatteschi, A. Grand, J. Laugier, L. Pardi, and P. Rey, Inorg. Chem. 27 , 1031, 1988. A. Caneschi, D. Gatteschi, J. Laugier, L. Pardi, P. Rey, and C. Zanchini, Inorg. Chem. 27, 2027, 1988. 249 85 86 87 88 89 90 91 92 93 94 95 A. Bencini, C. Benelli, D. Gatteschi, C. Zanchini, J. Am. Chem. Soc. 106, 5813, 1984. C. Benelli, D. Gatteschi, and C. Zanchini, Inorg. Chem. 23, 798, 1984. A. Caneschi, D. Gatteschi, J. Laugier, and P. Rey, "Organic and Inorganic Low Dimensional Crystalline Materials", P. Delhaes, M. Drilon, Eds.; Plenum Press, New York, London, p109, p389, 1987. A. Caneschi, D. Gatteschi, P. Rey, R. Sessoli, Inorg. Chem. 27, 1756, 1988, A. Caneschi, D. Gatteschi, J. P. Renard, P. Rey, adn R. Sessoli, J. Am. Chem. Soc. 111, 785, 1989. J. S. Miller, J. C. Calabrese, H. Rommelmann, S. R. Chittipeddi, J. H. Zhang, W. M. Reiff, and A. J. Esptein, J. Am. Chem. Soc. 109, 769, 1987, J. S. Miller and A. J. Epstein, J. Am. Chem. Soc. 109, 3850, 1987. J. S. Miller, Acc. Chem. Res., 21, 114, 1988. J. S. Miller, A. J. Epstein, W. M. Reiff, Chem. Rev., 88, 201, 1988. J. S. Miller, A. J. Epstein, and W. M. Reiff, Science,240, 40, 1988. A. Chakraborty, A. J. Epstein, W. N. Lawless, and J. S. Miller, Phys. Rev. B, 40, 11 422, 1989. J. S. Miller and A. J. Epstein, Phil Trans. R. Soc. Lond. A , 330, 205, 1990. C. Kollmar, O. Kahn, J. Am. Chem. Soc. 113, 7987, 1991, C. Kollmar, M. Couty, and O. Kahn, J. Am. Chem. Soc. 113, 7994, 1991. C. Kollmar and O. Kahn, Acc. Chem. Res. 26, 259, 1993. , C. M. Hurd, Contemporary Physics 23, 469, 1982. K. Mukai, H. Nishiguchi, and Y. Deguchi, J. Phys. Soc. Japan, 23, 125, 1967, K. Awaga, T. Sugano, and M. Kinoshita, Solid State Comm. 57, 453, 1986. K. Awaga, T. Sugano, and M. Kinoshita, Chem. Phys. Lett. 128, 587, 1986. K. Awaga, T. Sugano, and M. Kinoshita, J. Chem. Phys. 85, 2211, 1986. K. Awaga, T. Sugano, and M. Kinoshita, Chem. Phys. Lett. 141, 540, 1987. M. Kinoshita, Mol. Cryst. Liq. Cryst. 176, 163, 1989. D. E. Williams, Mol. Phys. 16, 145, 1969. K. M. Chi, J. C. Calabrese, J. S. Miller, adn S. I. Khan, Mol. Cryst. Liq. Cryst. 176, 185, 1989. K. Mukai, Bull. Chem. Soc. Jpn. 42, 40, 1969. K. Awaga and Y. Maruyama, Chem. Phys. Lett. 158, 556, 1989. K. Awaga, T. Inabe, U. Nagashima, and Y. Maruyama, J. Chem. Soc. Chem. Comm. 1617, 1989. M. Kinoshita, P. Turek, M. Tamura, K. Nozawa, D. Shiomi, Y. Nakazawa, M. Ishikawa, M. Takahashi, K. Awaga, T. Inabe, and Y. Maruyama, Chem. Lett. 1225, 1991. P. Turek, K. Nozawa, D. Shiomi, K. Awaga, T. Inabe, Y. Maruyama, and M. Kinoshita. Chem. Phys. 180, 327, 1991. L. Pauling "The Nature of the Chemical Bond", Cornell University Press, New York, p343, 1960. J. W. Linnett, J. Am. Chem. Soc. 83, 2643, 1961. 250 96 97 98 99 100 101 102 103 104 105 106 107 108 Y. Y. Lim and R. S. Drago, J. Am. Chem. Soc. 93, 891, 1970. A. M. Vasserman and A. L. Buchachenko, J. Struct. Chem. 7, 633, 1966. T. Sugano. M. Tamura, M. Kinoshita, Y. Sakai, and. Y. Ohashi, Chem. Phys. Lett. 200, 235, 1922. H. Iwamura, 197th National Meeting of the American Chemical Society, Dallas, TX, April 9—14, 1989. W. T. Borden, Diradicals; W. T. Borden, Ed.; Wiley, New York, 1982, W. T. Borden and E. R. Davidson, J. Am. Chem. Soc. 99, 4587, 1977. E. E. Seeger, P. M. Lahti, A. R. Rossi, and J. A. Berson, J. Am. Chem. Soc. 108, 1251, 1986. S. Kato, K. Morokuma, D. Feller, B. Davidson, W. T. Borden, J. Am. Chem. Soc. 105, 1791, 1983. B. B. Wright, M. S. Platz, J. Am. Chem. Soc. 105, 628, 1983. J. L. Goodman, J. A. Berson J. Am. Chem. Soc. 106 , 1867, 1984, M Rule, A. Matlin, D. E. Seeger, E. F. Hilinski, and D. A. Dougherty Tetrahedron 38, 787, 1982. T. Matsumoto, T. Ishida, N. Koga, and H. Iwamura. J. Am. Chem. Soc. 114, 9952, 1992. A. J. Novak, R. Jain, and D. A. Dougherty, J. Am. Chem. Soc. 111, 7618, 1989. J. A. Berson, Acc. Chem. Res. 11, 446, 1978, P. Dowd., Acc. Chem. Res. 5, 242, 1972, R. Jain, M. B. Sponsler, F. D. Coms, D. A. Dougherty, J. Am. Chem. Soc. 110, 1356, 1988. J. A. Berson, Acc. Chem. Res. 11, 446, 1978. M. Ballester, J. Castaner, J. Riera, A. Ibanez, and J. Pujadas, J. Org. Chem. 47, 259, 1982. M. Ballester, J. Riera, J. Castaner, A. Rodriguez, C. Rovira, and J. Veciana, J. Org. Chem. 47 , 4498, 1982. M. Ballester, J. Castaner, J. Riera, J. Pujadas, O. Armet, C. Onrubia, and J. A. Rio, J. Org. Chem. 49, 770, 1984. M. Ballester, Acc. Chem. Res. 18, 380, 1985. J. Veciana, A. D. Martinez and O. Armet, Rev. Chem. Intermediates, 10, 35, 1988, J. Veciana and A. Duran. J. Inclus. Pheno. 5, 173, 1987. O. Armet, J. Veciana, C. Rovira, J. Riera, J. Castaner, E. Molins, J. Rius, C. Miravitlles, S. Olivella, and J. Brichfeus. J. Phys. Chem. 91, 5608, 1987. J. Veciana, C. Rovira, M. I. Crespo, O. Armet, V. M. Domingo, and F. Palacio, J. Am. Chem. Soc. 113, 2552, 1991. J. Carilla, L. Julia, J. Riera, E. Brillas, J. A. Garrido, A. Labarta, and R. Alcala. J. Am. Chem. Soc. 113, 8281, 1991, J. Veciana, C. Rovira, N. Ventosa, M. I. Crespo, and F. Palacio J. Am. Chem. Soc. 115, 57, 1993. A. Rajca, J. Am. Chem. Soc. 112, 5889, 1990, A. Rajca, J. Org. Chem. 56, 2557, 1991, A. Rajca, S. Utamapanya, and J. Xu, J. Am. Chem. Soc. 113, 251 109 110 111 112 113 114 115 116 117 118 119 120 9235, 1991. S. Utamapanya and A. Rajca, J. Am. Chem. Soc. 113, 9242, 1991., A. Rajca, S. Utamapanya, and S. Thayumanavan J. Am. Chem. Soc. 114, 1884, 1992. Crystal Structure and Magnetic Susceptibility of a Hydrocarbon Free Radical: Tris (3,5-di-t-butylphenyl)methyl B. Kahr, D. V. Engen, and P. Gopalan, submitted for publication. A. Izuoka, S. Murata, T. Sugawara, and H. Iwamura, J. Am. Chem. Soc. 107, 1786, 1985., H. Iwamura, Pure & Appl. Chem. 58, 187, 1986. A. Izuoka, S. Murata, T. Sugawara, and H. Iwamura, J. Am. Chem. Soc. 109, 2631, 1987. K. Itoh, Chem. Phys. Lett. 1, 235, 1967., Y. Teki, T. Takui, K. Itoh, H. Iwamura, and K. Kobayashi J. Am. Chem. Soc. 108, 2147, 1986., K. Itoh, T. Takui, Y. Teki, and Takamasa Mol. Cryst. Liq. Cryst.176, 49, 1989. N. Nakamura, K. Inoue, H. Iwamura, T. Fujioka, and Y. Sawaki, J. Am. Chem. Soc. 114, 1484, 1992. H. Iwamura and N. Koga, Acc. Chem. Res. 26, 346, 1993. S. Murata and H. Iwamura, J. Am. Chem. Soc. 113, 5547, 1991. K. Inoue, N. Koga, and H. Iwamura, J. Am. Chem. Soc. 113, 9803, 1991. C. Ling, M. Minato, P. M. Lahti, and H. van Willigen, J. Am. Chem. Soc. 114, 9959, 1992. M. Kato, H. B. Jonassen, and J. C. Fanning, Chem. Rev. 64. 99, 1964. G. J. Sloan and W. R. Vaughan J. Org. Chem. 22, 750, 1957, G. J. Sloan and W. R. Vaughan J. Chem. Phys. 25, 697, 1956. E. G. Rozantsev, V. A. Golubev, M. B. Neiman, adn Y. V. Kokhanov, Bull. Acad. Sci. USSR. 559, 1965. E. G. Rozantsev, and V. A. Golubev, Bull. Acad. Sci. USSR. 695, 1965. R. M. Dupeyre, H. Lemaire, and A. Rassat J. Am. Chem. Soc. 87, 3771, 1965. E. G. Rozantsev, V. A. Golubev and V. A. Golubev Bull. Acad. Sci. USSR. 382, 1965. D. F. Evans, J. Chem. Soc., 2003, 1959. D. G. B. Boocock, R. Darcy and E. F. Ullman, J. Am. Chem. Soc. 90, 5445, 1968, D. G. B. Boocock and E. F. Ullman J. Am. Chem. Soc. 90, 6873, 1968. E. F. Ullman and D. G. B. Boocock Chem. Comm. 1161, 1969, P. W. Kopf, R. Kreilick, D.G. B. Boocock, and E. F. Ullman, J. Am. Chem. Soc. 92, 4531, 1970. F. Alies, D. Luneau, J. Laugier, and P. Rey, J. Phy. Chem. 97, 2922, 1993. P. M. Lahti and A. S. Ichimura Mol. Cryst. Liq. Cryst. 176, 125, 1989. O. H. Griffith and A. S. Waggoner, Acc. Chem. Res. 2, 17, 1969, H. Dugas Acc. Chem. Res. 10, 47, 1977. E. Ullman, J. H. Osiecki, D. G. B. Boocock. and R. Darcy J. Am. Chem. Soc. 94, 7049, 1972. 252 121 122 123 124 125 126 127 128 129 130 K. Mukai, T. Yano, and K. Ishizu, Tett. Lett. 22, 4661, 1981. K. Ishizu, H. Kohama, and K. Mukai, Chem. Lett. 227, 1978, K. Mukai, N. Iida, Y. Kumamoto, H. Kohama, and K. Ishizu, Chem. Lett. 613, 1980, K. Mukai, M. Yamashita, K. Ueda, K. Tajima, and K. Ishizu, J. Phys. Chem. 87, 1338, 1983. K. Mukai, M. Tanii, Y. Yurugi, K. Tajima, and K. Ishizu Bull. Chem. Soc. Jpn., 58, 322, 1985. P. J. Hay, J. C. Thibeault, and R. Hoffmann, J. Am. Chem. Soc. , 97, 4884, 1975. R. Hoffmann, Acc. Chem. Res. , 4, 1, 1971, R. Gleiter, Angew. Chem. , 86, 770, 1974. O. Kahn, Angew. Chem. Int. Ed. Engl. 24. 834. 1985. J. B. Goodenough, ” Magnetism and the Chemical Bond" , Interscience, New York, N.Y., 1963. J. Kanamori, Phys. Chem. Solid, 10, 87, 1959, P. W. Anderson, Solid State Phys., 14, 99, 1963. See also perchlorotriphenylmethyl: M. Ballester, J. Riera, J. Castaner, C. Badia, M.J. Monsé, J. Am. Chem. Soc. , 93, 2215, 1971. The ferromagnetic coupling of a pair of nitroxyl substituted benzo-lS-crown-S- ethers was accomplished with potassium ions. See: K. Ishizu, H. Kohama, K, Mukai, Chemistry Letters, 227, 1978. See also: K. Mukai, N. Iida, Y. Kumamoto, H. Kohama, K. Ishizu, Chem. Lett. , 613 1980; K. Mukai, M. Yamashita, K. Ueda, K. Tajima, K. Ishizu, J. Phys. Chem. 87, 1338 , 1983.; J. F. W. Keana, J. Cuomo, L. Laszlé, S. E. Seyerdrezai, J. Org. Chem. 48, 2647, 1983; K. Mukai, M. Tanii, Y. Yurugi, K. Tajima, Bull. Chem. Soc. Jpn. 58, 322, 1985. In related work, glasses of t1is(biphenyly1)methy1 gave triplet coupling, presumably due to stacking of radicals pairwise to form D3 n-complexes. See: W. Broser, H. Kurreck, W. Niemeier, Tetrahedron, 32, 1183, 1976. For a review of one-dimensional ferromagnetism in inorganic solids see: R. D. Willett, R. M. Gaura, C. P. Landee in J. S. Miller (Ed.) Extended Linear Chain Compounds, V. 3, Plenum, New York, 1983; p143. C. Kollmar and O. Kahn Acc. Chem. Res. 26, 259, 1993. K. Nath, P. L. Taylor, Mol. Cryst. Lig. Cryst., 180B, 389 , 1990.; K. Yamaguchi, H. Namimoto, T. Fueno, Mol. Cryst. Lig. Cryst., 176, 151 1989.; H. Iwamura, S. Murata, Mol.Cryst. Liq. Cryst., 176, 33 , 1989.; J. Veciana, C. Rovira, O. Armet, V. M. Domingo, M. I. Crespo, F. Palacio, Mol. Cryst. Liq. Cryst., 176, 77, 1989.; D. J. Klein, S. A. Alexander, M. Randic, Mol. Cryst. Liq. Cryst., 176, 109 1989. T. Hughbanks, M. Kertesz, Mol. Cryst. Liq. Cryst., 176, 115 , 1989.; J. S. Miller, A. J. Epstein, Mol. Cryst. Liq. Cryst., 176, 347 , 1989. 253 131 132 133 134 135 136 137 138 139 140 141 142 143 144 H. M. McConnell, J. Chem. Phys., 24, 764, 1956., D. J. E. Ingram, "Free Radical as Studied by Electron Spin Resonance,” Butterworths, p.111, 1958. H. Jukeikis and D. Kivelson J. Am. Chem. Soc. 84, 1132, 1962. O. Armet, J. Veciana, C. Rovira, J. Riera, J. Castaner, E. Molins, J. Rius, C. Maravitlles, S. Olivella, and J. Brichfeus J. Phys. Chem. 91, 5608, 1987. A. G. de Mesquita, C. M. Gillarvy, and E. Eriks, Acta Cryst. 18, 437, 1965; P. Andersen and B. Klewe Acta Chem. Scand. 21, 2599, 1967; A. Port, M. M. Olmstead, and P. P. Power J. Am. Chem. Soc. 107, 2174, 1985; J. Veciana, J. Riera, J. Castaner, and N. Ferret, J. Organomet. Chem. 297, 131, 1985. K.M. Mislow, Acc. Chem. Res. 9, 26 (1976). M. Ballester, J. Riera, J. Castaner, C. Badia, and J. M. Monso, J. Am. Chem. Soc. 93, 2215, 1971. K. S. Hayes, M. Nagamo, J. F. Blount, and K. Mislow, J. Am. Chem. Soc. 102, 2773, 1980. "Comparison of Twists in Isosteric Propellers: X-ray Structures of Tris(2,6- dimethoxy-phenyl)borane, Tris(2,6-dimethoxyphenyl)methyl cation, and Tris(2,6- dimethoxyphenyl)methyl Radical”; B. E. Kahr, J. E. Jackson, D. Ward, S. -H. Jang, and J. Blount, Acta Cryst., B48, 324, 1992. "Aryl Ring Twists in Tris(2,6-dimethoxyphenyl)-Z Tripod Ethers: Why is the Radical Different ?" S. J. Stout, P. Gopalan, J. E. Jackson, and B. Kahr, Submitted for publication. M. M. Olmstead and P. E. Power, J. Am. Chem. Soc. 108, 4235, 1986. K. Ishizu, K. Mukai, A. Shibayama, and K. Kondo, Bull. Chem. Soc. Jpn. 50, 2269, 1977. R. Terrence, O. Toole, N. Janet, B. Younathan, P. Sullivan, and T. J. Meyer Inorg. Chem. 28, 3923, 1989. See F. A. Cotton, G. Wilkinson, "Advanced Inorganic Chemistry" Fifth Ed., Wiley, NY. 1988, pp. 1288 ff., 1388 ff. T. P. Radhakrishnan, Z. G. Soos, H. Endres, L. J. Azevado, J. Chem. Phys. 85, 1126 (1986). H. Volz, W. Lotsch, Tetrahedron Lett., 27, 2275, 1969. S. Bank, C. L. Ehrlich, J. A. Zubieta, J. Org. Chem., 44, 1454, 1979. S. Bank, C. L. Ehrlich, M. Mazur, J. A. Zubieta, J. Org. Chem, 46, 1243, 1981. For a general overview, see "Organic Electrochemistry," 2nd Ed., M. M. Baizer, H. Lund, Eds. Marcel Dekker, New York, 1983, Chap. 2. R. W. Brockman and D. E. Pearson, J. Am. Chem. Soc. 74, 4128, 1952, P. Zuman, Chem. Listy, 47, 1234, 1953. I. R. Fox, R. W. Taft, Jr., and I. C. Lewis, J. Am. Chem. Soc. 81, 5343, 1959. 'lll 254 145 146 147 148 149 150 151 152 153 154 155 156 157 T. Berzins and P. Delahay, J. Am. Chem. Soc. 75, 5716, 1953. W. H. Reinmuth, L. B. Rogers and L. E. I. Hummelstedt, J. Am. Chem. Soc. 81, 2947, 1959. C. D. Ritchie, W. F. Sager Prog. Phys. Org. Chem. 2, 323, 1964. C. G. Swain, S. H. Unger, N. R. Rosenquist, and M. S. Swain, J. Am. Chem. Soc. 105, 492, 1983, C. Hansch, A. Leo, S. Unger, K. H. Kim, D. Nikaitani, and E. Liem, J. Med. Chem. 16, 1207, 1973, R. Popielarz and D. R. Arnold, J. Am. Chem. Soc. 112, 3068, 1990. C. D. Ritchie, "Physical Organic Chemistry", Marcel Dekker, New York, 102, 1975. T. H. Lowry and K. S. Richardson, ”Mechanism and Theory in Organic Chemistry", Harper & Row, New York, 149, 1987. J. J. Brooks and G. D. Stucky J. Am. Chem. Soc. 94, 7333, 1972. M. M. Kreevoy and D. C. Johnston, Croatica Chemica Acta 45, 511, 1973. D. Masilamani and M. M. Rogic J. Org. Chem. 46, 4486, 1981, W. D. Watson, J. Org. Chem. 50, 2145, 1985, U. Wriede, M. Fernandez, K. F. West, D. Harcourt, and H. W. Moore, J. Org. Chem. 52, 4485, 1987. E. Bay, D. A. Bak, P. E. Timony, and A. Leone-Bay, J. Org. Chem. 55, 3415, 1990. E. Iiehlmann and R. W. Lauener Can. J. Chem. 67 , 335, 1989. R. A. Eade, F. J. McDonald and H. P. Pham, Asust J. Chem. 31, 2699, 1978. D. C. Schlegel, C. D. Tipton, and K. L. Rinehart, Jr. J. Org. Chem. 35, 849, 1970, T. L. Davis and V. F. Harrington J. Am. Chem. Soc. 56, 129, 1934. H. Jukeikis and D. Kivelson J. Am. Chem. Soc. 84, 1132, 1962. K. Ishizu, K. Mukai, A. Shibayama, and K. Kondo, Bull. Chem. Soc. Jpn. 50, 2269, 1977. A. Rieker and H. Kessler, Tett. Lett. 1227, 1969; H. Kessler, A. Moosmayer, and A. Rieker, Tetrahedron, 25, 287, 1969. H. S. Gutowsky, C. H. Holm, J. Chem. Phys. 25, 1228, 1956., D. Kost, E. H. Carlson, and M. Raban, J. Chem. Soc. Chem. Comm. 656, 1971, J. Sandstrom, Dynamic NMR Spectroscopy; Academic Press, New York, 97, 1982. H. Freibolin, Basic One- and T wo- Dimensional NMR Spectroscopy, VCH, New York, 263, 1991. Pulsed EPR studies of Ion Binding in a Double-faced Paramagnetic IonOphore: Tris(2,6-dimethoxyethoxy)phenyl)methyl Radical S. —H. Jang, H. —1. Lee, J. MaCracken, and J. E. Jackson. Accepted for publication in. J. Am. Chem. Soc. H. G. Lohr and F. Vogtle Acc. Chem. Res. 18. 65, 1985. L. Kevan, in "Time Domain Electron Spin Resonance” L. Kevan & R. N. Schwartz, edit. Chapter 8, Wiley-Interscience, New York, 1979. 255 158 159 160 161 162 163 164 165 166 167 168 169 170 171 J. MaCracken, D.-H. Shin, and J. L. Dye, Appl. Magn. Reson. 3, 305, 1992. A. A. Shuben and S. A. Dikanov, J. Magn. Reson. 52, 1, 1983. W. Weltner, "Magnetic Atoms and Molecules”, Dover, New York, 1983. M. M. Baizer and H. Lund. ”Organic Electrochemistry", 2nd Edition, Chapter 2, Marcel Dekker, New York, 1983. S. Bank, C. L. Ehrlich, M. Mazur, and J. A. Zubieta J. Org. Chem, 46, 1243, 1981. G. Kothe, W. Summermann, H. Baumgartel and H. Zimmermann, Tett. Lett. 2185, 1969. S. Gauthier and J. M. Fréchet Synthesis , 383, 1987. See for example F. A. Cotton, G. Wilkinson, "Advanced Inorganic Chemistry" 5th Ed., Wiley Interscience, NY, 1988, p. 1385. J. E. Huheey, "Inorganic Chemistry; Principles of structure and reactivity" 2nd E. Harper & Row, NY, 1978, p71. "PCModel Molecular Modeling Software", K. E. Gilbert, J. J. Gajewski, M. M. Midland; Serena Software, Bloomington, IN 47402, (1987). There have been a number of reports of organic ferromagnetism. The more substantial accounts are those of the well-characterized crystalline metallocene derivatives cited. For other examples of purported organic ferromagnets in less well—characterized polymeric materials see: H. Tanaka, K. Tokuyama, T. Sato, T. Ota, Chemistry Letters, 1813, 1990 ; J. B. Torrance, S. Oostra, A. Nazzal Syn. Met. 19, 708 ,1987 ; Y. V. Korshak, T. V. Medvedeva, A. A. Ovchinnikov, V. N. Specktor, Nature, 326, 370 1987 ; Y. Cao, P. Wang, Z. Hu, S. K. Zhang, J. Zhao, Syn. Met. 27, B625, 1988 ; M. Ota; S. Otani; K. Kobayashi, Chem. Lett. 1179 , 1989. P. Grobet, L. V. Gerven, A. V. D. Bosch, and J. Vansummeren, Physica 86-868, 1132, 1977. W. Boon and L. V. Gerven, Physica B 155, 437, 1989. We have been unable to obtain any commercially available simulation programs for multiradical CW -EPR spectra. A careful structure determination with a review of the history of Chichibabin's hydrocarbon appeared recently: Montgomery, J. C. Huffman, E. A. Jurczak, M. P. Grendze, J. Am. Chem. Soc. 108, 6004, 1986. Although the singlet—triplet splitting in Chichibabin’s hydrocarbon has not been unambiguously determined, it is clear that the (singlet) wavefunction of the crystalline material studied by Montgomery has a significant biradical component. G. A. Jeffrey and W. Saenger in "Hydrogen Bonding in Biological Structures ” Springer-Verlag, Berlin Heidelberg New York, 1991 and references theirin. J. Herbich, J. Karpikuk, Z. R. Grabowski, N. Tamai, and K. Yoshihara J. Luminescence 54, 165, 1992. H. Tamiaki and K. Maruyama J. Chem. Soc. 256 Perkin. Trans. 2431, 1992. C. Turro, C. K. Chang, G. E. Leroi, R. I. Cukier, and D. G. Nocera, J. Am. Chem. Soc. 114, 4013, 1992. A. Harriman, Y. Kubo, J. L. Sessler, J. Am. Chem. Soc. 114, 388, 1992. D. N. Beratan, J. N. Betts, and J. N. Onuchic Science, 252, 1285, 1991. Y. Aoyama, M. Asakawa, Y. Matsui, and H. Ogoshi, J. Am. Chem. Soc. 113, 6233, 1991. 172 R. Hoffmann Acc. Chem. Res. 4, 1, 1971, P. J. Hay, J. C. Thibeault, and R. Hoffmann, J. Am. Chem. Soc. 97, 4884, 1975. R. Gleiter, Angew. Chem. 86, 770. 1974. {Erilitx LIBRQRIES IHHHI l I l STQTE U IV I l .4}. :x». .1... Fulani 1.: . .: i333 F2529“ :3