- . OVERDUE FINES: -. 25¢ per day per in- 7 . mgmm uinmv. MATERIALS: ' Place in book return to reiove . charge from aimiauon records' ELECTROCHEMICAL STUDIES OF SOME TRANSITION METAL CRYPTATES AND AN INVESTIGATION OF THE REACTION ENTROPIES OF TRANSITION METAL REDOX COUPLES By Edmund L. Yes A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Chemistry 1980 ABSTRACT ELECTROCHEMICAL STUDIES OF SOME TRANSITION METAL CRYPTATES AND AN INVESTIGATION OF THE REACTION ENTROPIES OF TRANSITION METAL REDOX COUPLES By Edmund L. Yee Electrochemical methods were used to monitor the aqueous complexation and redox chemistry of the trivalent and divalent europium and ytterbium (2.2.1) and (2.2.2) cryptates. All of these complexes were electrochemically reversible and substitutionally inert on the cyclic-voltammetric time scale. The substantially lower thermodynamic stabilities of the trivalent cryptates (com— pared to the corresponding divalent complexes) arise from large enthalpic destabilizations which outweigh smaller entropic stabili- zations. These differences can be explained in terms of the relative hydration of the uncomplexed trivalent and divalent ions and comply with the trends observed for alkali metal and alkaline earth cryp- tates having comparable ionic radii. The strong associations of the trivalent Eu and Yb cryptates with hydroxide and fluoride were found to be comparable to those reported for the aquated cations and these anions. For cations of approximately constant size, striking decreases in the cryptate formation and dissociation rates are observed as the cation charge increases from one to three. These rate varia- tions and their enthalpic and entropic components reflect the Edmund L. Yee increased changes in cation hydration which accompany cryptate formation when the cation charge increases. The dissociation of the lanthanide cryptates is substantially catalyzed by acid, and this can be attributed to the need to alter the ligand conformation before or during the removal of the cation from the cryptate cavity. The homogeneous self-exchange rate constants of Eu(2.2.l)3+/2+ and Eu(2.2.2)3+/2+ were calculated by using the Marcus cross-relation and rate data from cross reactions involving these cryptates. The greatly enhanced self-exchange rates of the cryptates relative to that for Hui.”2+ were ascribed to smaller solvent- and ligand- reorganization components in the activation barriers of the cryptate systems. The half-cell reaction entropies As;c, for some M(III)/(II) redox couples in aqueous solution were determined with a non- isothermal cell arrangement. Reaction entropies in the 36-49 e.u. range were found for aquo couples. Several substitutionally inert ruthenium (III)/(II) complexes were employed to scrutinize ligand effects on A8;c. The sizable decreases in A8;c which accompany the replacement of aquo ligands by ammines were attributed to less ex- tensive solvent-ligand hydrogen bonding by the latter ligands. Large reductions in As;c resulted when aquo and ammine ligands were replaced by anionic ligands. Examinations of couples containing chelating ligands disclosed substantial differences in Aséc between Co(III)/(II) couples and the corresponding Ru(III)/(II) and Fe(III)/(II) systems. These could be attributed to ligand- conformation and electron-delocalization factors. Edmund L. Yee A correlation between the self-exchange rate of a redox couple and its reaction entropy was found. This relation arises since both A8;c and outer- and inner-shell reorganization energies are related to solvent-structuring effects. Reaction entrOpies and formal potentials were employed to calculate the free-energy, enthalpic, and entrOpic driving forces necessary for a thorough comparison of the predictions of the con- ventional Marcus model of outer-sphere electron transfer with the experimental rate constants and activation parameters for cross reactions involving cationic transition metal complexes. The con- siderable disagreements between the experhmental and predicted values found at very large driving forces were compatible with the presence of sizable, unfavorable work terms in the formation of the binuclear collision complex. An analysis of the activation parame- ters indicated that these work terms may be due to the reorganiza- tion of solvating water molecules that is needed in order to form the highly charged collision complex. ACKNOWLEDGMENTS The author would like to express his appreciation to all the peOple, big and small, who helped him travel the long and winding road to the doctoral degree. The contributions of the following individuals deserve special mention: Dr. Michael Weaver has provided invaluable advice, guidance, and support. His insights have been shafts of gold amid the Stygian darkness of scientific uncertainty. My colleagues in the Weaver group have come through like pros in giving me some much-needed assistance in the laboratory. MOre importantly, they have been good and true friends. Special thanks go to Rose Manlove for transforming a scrawled, unreadable mass of dubious prose into a well—typed, legible mass of dubious prose. The staff of the Department of Chemistry has been an important factor in much of my progress as a graduate student. The Air Force and Navy have supported the author through a good deal of his graduate career. One can only marvel at the manner in which money is expended in the never-ending struggle against the spread of communism. Financial support by the Department of Chemistry and the General Electric Foundation is also gratefully acknowledged. ii iii The encouragement given to me by my brothers and sisters is profoundly appreciated. MOst importantly, I would like to thank my parents, without whom this all would have been impossible. TABLE OF CONTENTS LI S T OF T“ LES O O O O O O O O O O O O C O O O C O O O 0 LIST OF FIGURES. . . . . . . . . . . . . . . . . . . . . INTRODUCTION . CHAPTER I CHAPTER II HISTORICAL BACKGROUND. . . . . . . . . . . A. Historical Background - Cryptates. B. Reaction Entropies . . . . . . . . . . C. Homogeneous Electron-Transfer Kinetics of Transition Metal Complexes . . . . EXPERIMENTAL ASPECTS . . . . . . . . . . . A. Apparatus. . . . . . . . . . . . . . . B. Electrochemical Techniques . . . . . . 1. Common Features. . . . . . . . . . 2. Cyclic Voltammetry . . . . . . . . 3. Polarography . . . . . . . . . . . 4. Preparative Electrolyses . . . S. Potentiometric Methods . . . . . . iv Page viii xii 16 20 33 34 34 34 35 36 36 37 CHAPTER III C. D. Materials . . . . . . . . . . . . . . . 1. 2. 3. Cryptates . . . . . . . . . Reagents and Solvents . . . . Transition Metal Complexes and Compounds . . . . . . . . . . . Kinetic Methods . . . . . . . . . . . . 1. 2. 3. Cryptate Aquation Studies . Homogeneous Redox Reactions Princ iples O O O O O O O O O O O O O Homogeneous Redox Reactions Experimental Specifics. . . . . . ELECTROCHEMICAL STUDIES OF EUROPIUM AND YTTERBIUM CRYPTATE FORMATION IN AQUEOUS SOLUTION. O O O O O I O O O O O O O O O O O A. B. IntrOduc t ion 0 O O O O O O O O O O O O O Res‘llt s O O O O O O O O O O O O C O O O 1. 2. Complexation Thermodynamics . . . . :Conplexation and Dissociation Kinet ics O O O O O O O O O O O O O O DisCUSSiono O I O O O O O O O O O O O O 1. 2. The Effects of varying the Cation Charge upon Cryptate Thermodynamics . Association of Lanthanide Cryptates Wit h smal 1 An ion 8 O O O O O O O O O The Effects of Varying the Cation Charge upon Cryptate Substitution Kinetics O O O O O O O O I O O O O O 37 . 37 38 38 42 42 43 45 48 49 49 49 63 67 67 74 . 7S CHAPTER IV CHAPTER V vi REACTION ENTROPIES OF TRANSITION METAL Rwa COUPLES O O O O O O O O O O O I O 0 A. Introduction. . . . . . . . . . . . B. Determination of Reaction Entropies . C. Practical Aspects of Nonisothermal cells C O I C O O O O O O O O O O O O D. Results . . . . . . . . . . . . . . . 1. Aqua Couples. . . . . . . . . 2 . Mthenium(III)/ (II) and Osmium(III)/(II) Couples Containing Mbnodentate Ligands . . . . . . . . . 3. Redox Couples Containing Chelating Ligands . . . . . . . . . . Discussion. . . . . . . . . . . . . . . . 1. Ligand Effects on Reaction Entropies. 2. Correlations between Outer-Sphere Self-Exchange Rates and Reaction Entropies . . . . . . . . . . . . ACTIVATION PARAMETERS FOR HOMOGENEOUS OUTERPSPHERE ELECTRON-TRANSFER REACTIONS: COMPARISONS BETWEEN SELF-EXCHANGE AND CROSS REACTIONS USING THE MARCUS THEORY . A. Introduction. . . . . . . . . . . B. Kinetic Formulations and Results. . . 1. Free Energies of Activation . . 2. EntrOpies and Enthalpies of Activation. . . . . . . . . . . C. Discussion. . . . . . . . . . . . . . .102 .106 .106 .119 .133 .134 .136 .136 .156 .173 vii CHAPTER VI HOMOGENEOUS REDOX KINETICS OF THE EUROP I W cm TATES O I O O O I O O O O O O O O A. B. C. Introduction. . . . . . . . . . . . . . . Resu1ts O O O O O O O O O O O O O O O O 0 Discussion. 0 O O O O O O O O O O O O O O 1. Reliability of the Self-Exchange Parameters of the Eu Cryptates. . . 2. Mechanistic Implications of the Inverse-Acid Dependence of k12' . . 3. The Effect of Cryptate Formation on the Intrinsic Redox Reactivity of the Eu(III)/(II) Couple. . . . . . CHAPTER VII MISCELLANEOUS EXPERIMENTS . . . . . . . . . . A. Electrochemistry of Other Transition Metal Cryptates . . . . ; . . . . . . . The T1(2.2.2)+/T1(Hg) Couple as a Probe of Cryptate Thermodynamics in Nonaqueous Solvents. . . . . . . . . . . . . . . . . CHAPTER VIII CONCLUSIONS AND SUGGESTIONS FOR FURTHER WORK. A. B. LIST OF REFERENCES . COnclusions . . . . . . . . . . . . . . . 1. Transition Metal Cryptates. . . . . . 2. Reaction Entropies. . . . . . . . . . Suggestions for Further WOrk. . . . . . . O O O O O O I O O O O O O O O O O O O O 180 . 181 182 190 190 191 197 202 203 207 214 215 215 . 217 220 224 LIST OF TABLES Table Page 1 Thermodynamics of some Europium and Ytterbium.Cryptate Redox Couples in NOncomplexing Media at 25°C. . . . . . . . . . . . . . . .54 2 Chmulative Formation Constants for Association between Europium and Ytterbium Cryptates and Fluoride and Hydroxide Ions . . . . . . . . . . . . . . . . . . . . . .62 3 Rate Constants and Activation Parameters for Dissociation of Europium and Ytterbium.Cryptates at 25°C. . . . . . . . . . . . . .64 4 Acid Acceleration of Europium Cryptate Dissociation at 25°C . . . . . . . . . . . . . . . . . . .66 5 Anion Effects on Eu(2.2.l)3+ Dissociation. . . . . . . . .67 6 Free Energies, Enthalpies, and EntrOpies of Complexation of various Metal Cations with (2.2.1) and (2.2.2) Cryptands at 25°C . . . . . . . .69 7 Rate Constants and Activation Parameters for Formation and Dissociation of some (2.2.1) Cryptates at 25°C. . . . . . . . . . . . . . . . .77 viii Rate Constants and Activation Parameters for Formation and Dissociation of some (2.2.2) Cryptates at 25°C. . . . . . . . . . . . . . . . . .79 Reaction Entropies for various M(OH2):+/2+ Redox Couples. . . . . . . . . . . . . . . . . . . . . . . .96 Reaction EntrOpies for Various Ru(III)/(II) and Os(III)/(II) Redox Couples Containing Unindentate Ligands. . . . . . . . . . . . . . . . . . . . .98 Reaction Entropies for various M(III)/(II) Couples Containing Chelating Ligands . . . . . . . . . . . 103 Reaction Entropies for various Redox Cbuples of the Type MLéLKCIII)/(II). . . . . . . . . . . . 113 Comparison between Activation Free Energies for Some Ruthenium(III)/(II) Self—Exchange . Reactions and the Reaction Entropies for the Corresponding Redox Couples at 25°C. . . . . . . . . . . . 122 Comparison between Kinetic Parameters for Selected Outer-Sphere Self-Exchange Reactions and Reaction Entropies for Corresponding Redox Couples at 25°C. . . . . . . . . . . . . . . . . . . 125 Kinetic and Thermodynamic Parameters of Some Self-Exchange Reactions at 25°C . . . . . . . . . . . 140 Kinetic Parameters of Some Self-Exchange Reactions at 25°C. . . . . . . . . . . . . . . . . . . . . 142 Driving Forces and Rate Constants for Cross Reactions Between Aquo Cations at 25°C. 0 O O O I O I O O C O O O O O O O O O O O O O O O 145 Free Energies of Activation and Driving Forces for Cross Reactions Between Aquo “tions at 25°C. 0 O O O O O O O O I O O O O O O O O O O O 147 Driving Forces and Rate Constants for Cross Reactions Involving NOn-Aquo Cations at 25°C. 0 C O O O O O O O O O O O O O C O O O O O O O O O 150 Free Energies of Activation and Driving FOrces for Cross Reactions Involving Non-Aqua cations at 25°C 0 o o o e o o o o e e o o o e o o 152 Thermodynamic Parameters for Some Cross Reactions at 25°C. . . . . . . . . . . . . . . . . . 159 Comparison of Activation Parameters for Cross Reactions Determined at 25°C with Those Calculated from Marcus Theory. . . . . . . . . . . . 161 Activation Parameters for some Cross Reactions Having Large Driving Forces. Effect of Including Entropic and Enthalpic Werk Terms on Marcus Predictions . . . . . . . . . . . . . 166 Rate Constants and Equilibrium Constants for Cross Reactions Involving Europium cryptates O O O O O O O O O O I O O O O O O O O O O O O I O 187 Self-Exchange Rate Constants for Selected NOncryptate Couples. . . . . . . . . . . . . . . . . . . . 188 26 27 .xi Self-Exchange Rate Constants for Eu(2.2.1)3+/2+- and Eu(2.2.2)3+/2+. . . . . . . . . . . . . 190 Formal Stability Constants (log Ks) for Univalent (2.2.2) Cryptates in Dimethylsulfoxide and Acetonitrile at 25°C. 0 o ‘0 o o o o e o e o o e o o o o o o o e o e e o 213 Figure LIST OF FIGURES Structures of Some Common Cryptands . . . . . . . . . . . 6 Cathodic-anodic cyclic voltammograms of BuzzandEu(2.2.l)3+...................52 Plot of Bi, the formal potential for the Eu(2.2.l)3+/2+ couple in the presence of fluoride anions, vs. the logarithm of the fluoride concentration. Ionic strength held constant at 0.5 with sodium perchlorate. . . . . . . 60 Plot of the activation energy for some outer- sphere self-exchange reactions vs. the reaction entropy for the redox couples . . . . . . . . . . . . . . 127 Experimental (AH: for cross reactions 2)corr corrected for Debye—Hfickel work terms plotted against (AHf2)calc. . . . . .l. . . . . . . . . . 170 Experimental (ASE for cross reactions 2)corr corrected for Debye-Hflckel work terms plotted against (A8i2)calc. . . . . . . . . . . . . . . . . . . . 172 xii INTRODUCTION Electron-transfer reactions that involve transition metal ions form the basis of many scientifically and technologically important processes, so it is crucial that attempts be made to reach a better understanding of the factors which govern the redox thermodynamics and kinetics of these cations. One effective method to scrutinize the behavior of a metal ion redox couple is to vary its ligand environment in some appropriate fashion. This variation is often accompanied by changes in the standard potential of the couple and in the lability of the respective oxidation states, by modifications in the degree and type of solvent structure around the metal complex, and by alterations in the mechanism by which electron transfer occurs. Among the numerous ligands and complexes which have been or could be employed for such studies are cryptands, a class of polyoxadiaza- macrobicycles, and cryptates, the complexes which these cryptands form with metal ions. In these complexes, the cation is encapsulated ‘Within a molecular cavity defined by the ligand, and this uncommon structure is expected to modify the redox prOperties of the enclosed Inetal ion couple in an interesting and unusual fashion. The elucida- tion of these ligand effects will require the investigations of the cryptate ligation processes and redox thermodynamics that are re- counted in Chapter III along with the studies of the redox kinetics of these complexes which are described in Chapter VI. While little work has been done on electroactive cryptates, some of the previous findings regarding cryptate complexation thermodynamics and kinetics do have a bearing on the results detailed in Chapter III. Therefore, this earlier material is summarized in Section A of Chapter I. ‘While some useful trends concerning ligand effects may be gleaned from the transition metal cryptates, their unusual structure may restrict the conclusions of this study to a number of specific points. One way to achieve a broader view of ligand influences is to determine the ionic entrOpies of transition metal ions and the factors which affect these quantities. Such measurements can be utilized to obtain information on the degree of hydration around an ion and on the extent and type of interactions between the ligands and the solvent. In Chapter IV, entrOpy data for a series of related transition metal couples are used to ascertain the ligand effects on the redox thermodynamics of these complexes. 1 Since solvent reorganization is a major component of the acti- vation barrier in electron transfer reactions, one would expect that the solvent structure around an ionic reactant would have a strong influence on the redox kinetics of this species. Ionic entrOpies should provide insights into this structure and thus will be helpful in explaining trends in redox reactivity. In fact, a correlation between entrOpies and redox rate parameters is found and discussed in Chapter IV. The entrOpies of some transition metal ions have been determined, but very little has been done in a systematic manner, especially with respect to the ligand dependences of these entropies. Section B of Chapter I attempts to introduce some unity to the field of ionic entropies. It is especially important to make some sense of the various scales of entropy that are employed since no standard con- vention has been agreed upon. The combination of the appropriate ionic entrOpies will yield the entrOpy of reaction for complexation and redox processes. The former application will prove useful in Chapter III when cryptate stabilities are discussed in entropic and enthalpic terms. The latter employment of entropies will be essential in Chapter V when the limitations of the contemporary theories of homogeneous electron- transfer kinetics are examined. As a preparation for this analysis, Section C of Chapter I briefly summarizes these theories. An over- view of some of the principal electron-transfer mechanisms is also given since these concepts and the theories mentioned above will be needed to analyze and to understand the kinetic data for cryptate redox reactions presented in Chapter VI. CHAPTER. I HISTORICAL BACKGROUND A. Historical Background - Cryptates Cryptands were first synthesized by Lehn and co-workers in 1969 (1,2). (The macrobicyclic ligands used in this study are more pre- cisely designated [2]-cryptands to distinguish them from the tri— cyclic [3]-cryptands and the tetracyclic [4]-cryptands.) All of these cryptands contain three polyether strands attached to two bridgehead nitrogens. The common names of the cryptands consist simply of the number of ether oxygens in each of these strands. Figure 1 gives the structure and the common names of the cryptands used in this work. The polyether strands of a cryptand define a three-dimensional molecular cavity. Thus, the cryptands (2.2.2), (2.2.1), and (2.1.1) have a cavity radii of 1.4, 1.1, and 0.8 A, respectively (3). A metal ion of the proper size can be encapsulated within this cavity to yield an inclusion complex (3). In actual practice, these cryptate complexes are frequently made simply by mixing the ligand and a suitable salt of the metal ion in an appropriate solvent. The crystal structures of a number of these cryptates have been determined, principally by Weiss and co-workers (4,5). The inclusion complexes have several structural features in common. The metal ion lies within the molecular cavity of the ligand. The cryptand has an endb-endb conformation; i.e., the lone pairs on the nitrogen atoms point into the cavity. It is also possible for the ligand to adopt configurations where both electron pairs are directed away from the cavity (exo-exo} or where one pair points into the cavity while the = l, n = 1 (2.2.2) m = l, n = 0 (2.2.1) = 0, n = 1 (2.1.1) Figure 1. Structures of some common cryptands. other points away (endb-exo) (3). When the size of the metal ion closely matches that of the cryptand cavity, the heteroatoms of the ligand tend to be symmetrically distributed around the ion. On the other hand, some distortion of the ligand occurs when the metal ion is too large or too small for the cryptand cavity (5). A good deal of the interest in and the work on cryptands has stemmed from the discovery that these ligands are capable of forming stable complexes with alkali metal cations in aqueous and nonaqueous solutions (3,6). Along with the crown cyclic polyethers first synthe- sized by Pederson (7-10), the cryptands gave rise to the new field of the study of alkali metal complexation. Investigations of the in- fluences on the stability of the alkali metal cryptates have found that each cryptand has its own size selectivity. Simply stated, for a given cryptand, the most stable alkali metal cryptate has the closest match between the radius of the cation and that of the ligand's molecular cavity. The other alkali metal complexes of this particular cryptand would be less stable and would contain a cation which was unsuitably large or small with respect to the ligand cavity. Thus Na+ with a radius of 1.02 A.(1l) forms the most stable alkali metal complex with the (2.2.1) cryptand (cavity radius - 1.1 A.(3)), while K+ (radius - 1.38.A) plays the corresponding role for (2.2.2), which has a cavity radius of 1.4 A.(3). However, it has been noted (12) that one cannot think of this selectivity as a simple consequence of one cation having a better steric fit into the ligand cavity. The observed selec- tivity is the product of a combination of several factors, including the change in enthalpy which results when the solvent molecules in the metal ion's first coordination sphere are replaced by the ligand and the entropic effects arising from the desolvation of the cation and cryptand which accompanies the formation of the complex (12) . Alkaline earth cations also form stable cryptates and several studies of the formation thermodynamics of these complexes have been published (3,6,12). For alkali metal and alkaline earth ions of comparable radius, the divalent ion has generally been found to form more stable cryptates (6,13). The greater stability of the divalent cryptates is not merely due to greater ion—dipole interactions between the cation and the bonding sites of the ligand. An examination of the complexation thermodynamics for alkali metal and alkaline earth cryptates (12) revealed that the enthalpies of complexation for the divalent ions were comparable to or less favorable than those for monovalent ions of similar size. Thus, the increased stabilities of the alkaline earth complexes arise from entropic factors. The hydra- tion around one of these divalent ions is more extensive than that for the corresponding monovalent ion, so one expects a more favorable entropy of complexation when the divalent hydration shell is renoved upon the encapsulation of the metal ion within the cavity of the cryptand. Studies of cryptate thermodynamics have not been limited to alkali Inetal and alkaline earth complexes. The formation constants for a n‘J-‘lnber of transition metal cryptates have been determined (14-17) with the same potentiometric methods which were used for most of the alkali 2+, P 2+ metal and alkaline earth systems. While Cd 13 a 382+: Ag+. and + T1 form a number of stable cryptates, the d-block transition metal 10118, Niz+, an+, and Coz+, do not show much tendency to follow suit. of the d-block ions, only Cu2+ forms cryptates with appreciable Q“ IAI DU. stability constants. The relative instability of the cryptates of the d—block ions has been explained as a combination of size incompati- bility and hard-soft acid-base considerations (15) . The d-block cations studied all have radii of about 0.70 to 0.75 A (18) which would be too small for the (2.2.1) and (2.2.2) cavities. The (2.2.1) cryptand has a cavity radius of 0.8 Awhich would be a good match for the cation radii given above. However, the (2.2.1) complexes are still not very stable. The majority of the donor groups of this ligand are OXygen atoms, but the d-block ions have a greater affinity for nitrogen donors. Therefore, the instability of these (2.1.1) complexes is at least partly the result of relatively unfavorable interactions between the cations and the donor atoms of the ligand. To verify this, Lehn and Montavon synthesized a ligand which was analogous to (2.1.1) except that the two oxygen atoms in the diether strand were replaced by two N—CH3 groups. This tetraaza ligand formed complexes with d-block ions which were at least five orders of magnitude more stable than the cor- responding (2.l.l) cryptates (15). While cryptate thermodynamics has received a good deal of attention, the kinetics of cryptate formation and dissociation has not been neglected either. Studies have been performed in aqueous and non- a><1ueous solution, and several kinetic techniques have been employed, ineluding proton and alkali metal NMR (nuclear magnetic resonance) (19-23), temperature jump (24), and stopped-flow (25-28). Investiga- t343113 of cryptate ligation kinetics have concentrated almost exclu- 31\iely on the alkali metal and alkaline earth complexes. (The only Q‘ertions to this pattern are T1(2.2.2)+ (19) and Ag(2.2.l)+ (27).) Ho"Never, much useful information may be gleaned from these studies. 10 What follows will be concerned with aqueous solutions since these are the media employed in the kinetic studies of the present work. Within a given charge and ligand type, cryptate stability is more clearly reflected in the dissociation rate constant kd rather than the formation rate constant kf. The most stable cryptate will generally have the smallest kd (26) although a number of anomalies (e.g. Sr(2.2.l) 24-) are present. A comparison between the k and k f d values of alkali metal and alkaline earth cryptates reveals that greater differences tend to be found between charge types than within a given charge type. The kf and kd for alkaline earth cryptates are consis- tently much smaller than those for the alkali metal complexes. Some insight into the factors which determine these relative values has been gained by examining the activation entropies of the formation and dissociation reactions. In particular, the activation entropies for the dissociation process usually are negative quantities having appreciable magnitudes. This has led a number of workers (13,26) to conclude that the activated complex for the formation and dissociation reactions bears a greater resemblance to the separated reactants (i.e., metal ion and ligand) than it does to the cryptate product. One can think of the dissociation reaction as proceeding via a transi- tion state in which some water molecules coordinate to the metal ion. while the metal ion is still bound to the cryptand, there is a good Possibility that the cation is no longer inside the molecular cavity (25). The alkaline earth ions, as a consequence of their greater charge, will tend to structure the solvent to a greater extent in this transi- tion state than alkali metal ions would. This is supported by a study Of the activation entropies for the dissociation of the (2.2.2) 11 cryptates. The Na+ and K+ complexes have positive activation entropies 2+ 2+ 2+ 9 while the corresponding quantities for the Ca Sr , and Ba cryptates are considerably more negative. Similar trends may be seen for the (2.2.1) complexes, although the data for the K+ complex are not available. It has also been noted (13,26) that the kf values for the alkali metal and alkaline earth cryptates follow the same pattern as the rates of water exchange in the hydration shells of the cations although the actual formation rates are several orders of magnitude slower than those for the water exchange. The correspondence between kf and the VPater exchange rates has led to the conclusion that partial or complete desolvation of the cation occurs prior to or during the rate-determining step of cryptate formation (26). Of course, the extent of desolvation which is thought to occur here must be in harmony with the general description of the activated complex given in the preceding paragraph. The much smaller kf values relative to the water exchange rates have been explained in terms of two alternative or possibly coexistent effects (26). To begin with, the concentration of the reactive con- formation of the cryptand may be quite low. This would imply an .unfavorable pre-equilibrium before the rate-determining step. In a~<1vziition, a large steric barrier to cryptate formation may exist as a. consequence of the possibility that first bond formation may be difficult with the relatively rigid cryptands. Under a given set of conditions, cryptate formation and dissocia- t10:)n follow simple second-order and f irst-order rate laws, respectively (26). However, kinetic studies of these processes are not completely free from complications. Cox and Schneider have found that the 12 dissociation rates of many cryptates are acid catalyzed (27). The experimental dissociation rate constants Rd had acid dependencies of the form k3 - kd + ka [3*] (1) The presence of the acid catalyses furnishes one with information on the mechanism of the dissociation reaction. Cryptates can be con- sidered to exist in solution as an equilibrium mixture of three conformations: the endo-endo, exo-endo, and exo-exo. The endo-endo is the most stable thermodynamic form and is the conformation found in the solid state, as mentioned previously. However, dissociation can occur when the cryptate is any of these three forms (19). Because at least one nitrogen in the exo-endo and exo—exo forms is directed away from the interior of the cavity, one can conceive of a mechanism in which protonation at this site occurs prior to the rate-determining. removal of the metal ion from the cryptate cavity (27) . Such a pre- ceding step could give rise to the 1:3 [H+] term in eqn (1). The kcl term can then be seen to stem from a mechanistic pathway in which the metal ion exits the cryptate cavity while the ligand remains in an endo-endo conformation. Because the lone pairs on the nitrogens renain directed into the cryptate cavity, there is little chance for Protonation to occur prior to the rate-determining removal of the Qation. The combination of the two pathways described here results in the two terms of eqn (1). The observed acid catalyses originate in the stabilization of an intermediate (the exo-endo or era-3x0 c(Intiformation) via the protonation of one of the nitrogens on the ligand. It is also enlightening to note that the dissociation rates of several 13 cryptates (e.g. Sr(2.2.1)2+) show little, if any, acid dependence. This has been interpreted as an indication that the conversion of the endo—endo to exo-endo and era-9x0 forms is the rate-determining step in the equation of these cryptates (27). Thus no protonation preceding the rate-controlling step is found for either the endo-endo or the exo-endo/exo—exo dissociation pathways. Several electrochemical studies of cryptates have been performed. The great majority of these have been carried out by Peter and Gross neith propylene carbonate as the solvent (29-36). Britz and Rnittel have produced the only electrochemical study conducted in aqueous solution (37). This work dealt with an investigation of the adsorption of K(2.2.2)+ at a dropping mercury electrode. A combination of drop- time measurements and a.c. polarography was employed to determine the maximum surface coverage and the diffusion coefficient of the K(2.2.2)+ ion. Most of the work of Peter and Gross has been devoted to the electro- chemistry of alkaline metal and alkaline earth cryptates in propylene carbonate (29-33,35). Several points have emerged. The electroreduc- tions of the cryptates occur at considerably more negative potentials than those for the corresponding solvated metal ions. This result is not unexpected. The oxidized form of the alkali metal or alkaline earth is stabilized by complexation with the cryptand. An electrode of greater reducing power (i.e., more negative potential) is required to reduce such a species to the metallic state. Prolonged electrolysis of solutions of K(2.2.2)+ has shown that the products of the electro- reduction of this complex are the amalgam K°(Hg) and the free ligand (35) . Gisselbrecht, Peter, and Gross have also studied the electrochemistry 14 of other cryptates. In cyclic voltammograms of the Tl(2.2.2)+ complex, one could observe peaks corresponding to the electroreduction and electrooxidation of the cryptate (34). In contrast, the voltammograms of the alkali metal and alkaline earth cryptates have no anodic peaks which can be attributed to an oxidation which involves the complex (32,35). The complexes of (2.2.1) and (2.2.2) with Fe2+, Coz+, and Ni2+ have been investigated as well (36). The electrochemical reductions of these (:ryptates in prOpylene carbonate were found to be irreversible deposi- tion reactions and to occur at'more negative potentials than the Ireductions of the corresponding solvated metal ions. (The Ni(2.2.l)2+ complex is the sole exception to this second trend.) While these alkali metal, alkaline earth, and transition metal cryptates which tindergo deposition reactions are interesting electrochemical systems for several reasons, they are far from ideal for studies of the funda- muental parameters which affect redox reactivity. For these complexes, the analysis of electron-transfer kinetics is complicated by the ligation reactions which are involved in the deposition process. Given the above considerations, one comes to the conclusion that a more suitable cryptatewould be one in which both halves of a metal ion redox couple would remain in complexed form and in solution. These Properties would yield mechanistically simpler systems which are free from the preceding or following chemical steps that accompany a deposition. These simple, straightforward systems would be a crucial c<>tnponent in a fundamental and unambiguous investigation of the factors ‘thich control redox thermodynamics and kinetics. One may examine the preceding information and trends and speculate 15 that a europium or ytterbium cryptate couple might have the desirable features which were described above. The radii of Euz+ (1.17 A (11)) and Yb2+ (0.93 A (18)) are such that one would expect these ions to have stable (2.2.1) and (2.2.2) complexes, if the size criteria for- mulated from the alkali metal and alkaline earth systems are valid for these lanthanide ions. It has been stated the chemical properties of divalent lanthanide ions are similar to those of the alkaline earth ions having comparable radii (38) . Thus, there is no reason to suspect that the factors which govern the thermodynamic and kinetic stabilities of the alkaline earth cryptates will be qualitatively different to those for the Eu2+ and Yb2+ cryptates. In aqueous solution, both of these divalent lanthanide ions can be converted to the trivalent ion in a simple one-electron oxidation. Whether a similar one-electron interconversion in aqueous solution can be found for the cryptates of Eu and Yb is one of the objectives of this investigation. In addition, the kinetic stabilities of the tri- valent and divalent forms of these complexes will be determined. This information would be useful in determining the suitability of the lanthanide systems for redox studies since complexes with slow dissoci- ation kinetics are required in order to minimize the mechanistic com- Plications arising from ligation reactions which coincide with the electron-transf er proc es s . Furthermore, a study of the thermodynamic stabilities of the Eu and Yb cryptates will be necessary. This will be needed for determin— ing the thermodynamic contributions to the redox reactivity of the Eu and Yb complexes. As usual, an analysis in terms of enthalpies and er1t:ropies will provide more detailed information than could be acquired 16 solely from free energies. One effective method for obtaining such entropic and enthalpic data is through the employment of ionic entrOpies and particularly reaction entropies as is now discussed. B. Reaction EntrOpies Partial molal entropies can be exceedingly useful when one investi- gates the thermodynamic properties of a species in solution. The partial molal entropy of the i th component of a system §i can be defined as the temperature dependence (under constant solution com- position and pressure) of Hi, the chemical potential of component 1 (3111/31)? - S1 (2) In solution studies one frequently focuses attention on. one ion. Under these circumstances, it would seem appropriate to think in terms of partial molal entrOpies for single ions. Strictly speaking, only the partial molal entropy of a neutral salt (i.e., the sum of the ionic entrOpies for the cation and anion) is experimentally accessible (39). However, methods are available for estimating single ion entropies. One may arbitrarily assign an entrOpy value to some ion and thus Obtain a relative scale of single ion partial molal entropies. It is conventional to use the scale based on the assumption that SE... - O (39). An alternative approach is to estimate the absolute entropy of an ion in solution by means of an extrathermodynamic assumption. Several different assumptions have yielded similar values of S12... (40). An cal deg-1 mol-l) was obtained approximate value of SE... - -5 e.u. (e.u. by each of these methods (40). Once one absolute ionic entropy is set, the relative scale of entropies mentioned above can be easily converted into an absolute scale. While the scope for interpretation is somewhat 17 smaller with the relative scale, an ionic entropy from either scale contains valuable information on the interactions between an ion and the solvent molecules surrounding it. Because transition metal ions are used extensively in studies of redox thermodynamics and kinetics, the entropies of such ions are of especial interest to the present work. Some investigations of these systems have already been conducted. The entropies of these ions have been found to depend markedly on the ionic charge and on the types of ligands coordinated to the metal ion (41-44). In examining the various possible reasons for this wide variation in entropies, one can see how electrostatic effects (45,46) contribute to the first dependence. The "structure-making" and "structure-breaking" ability of the ion in solution will also have a decided influence on the ionic entropy (47-49) . While individual ionic entropies have a great deal of utility, a variation of this concept is particularly appropriate when dealing with redox systems. One can consider 115:1, the difference in partial molal entropies of the reduced and oxidized forms of a redox couple. Under this definition AS]; is the reaction entropy for a half-cell reaction and can be regarded as the entrOpic analog to the half-cell electrode potential. One can employ 138;: in a variety of circumstances for a substantial number of purposes. A relative measure of the degree of Solvent ordering around the oxidized and reduced forms of a metal ion complex is available through the use of As;c. The entropy of reaction for a homogeneous redox reaction can be calculated by taking the dif- ference between the Asgc values of the two redox couples involved. I—-arge AS:c values are frequently found, so that this term can be the 18 predominant constituent of the free-energy driving force for an electrode reaction (50). By combining the appropriate values of As;c one can obtain the difference in the entrOpies of formation between the oxidized and reduced forms of a metal ion complex. This will definitely be applicable when comparing the complexation thermo- dynamics of the trivalent and divalent lanthanide cryptates. As evidenced by the above discussion, a knowledge of As;c values can be very useful. Unfortunately, a comparative shortage of data wists in the literature. Moreover, some of these data are contradic- t°tY . possibly as a consequence of the different entrOpy scales which are in use. Nonetheless, it has been determined that the A31; values °f transition metal couples are considerably influenced by the ligands which are coordinated to the metal ion. For one-electron aquo cation redox couples, Asa: values in the 40 to 50 e.u. range are found, pro- vIlded that one enploys the normal convention and writes the half-cell reaction as a reduction (41,45,49). In contrast, metal ion complexes cuntaining large organic ligands have been found to have much smaller 481°C values (51). This effect has been ascribed to the efficacy of the Organic ligands in shielding the cation from the surrounding water tllolecules (51). While some helpful points have emerged from the results described above, it is clear that a study of wider scape and greater depth is 1leaded. In particular, Asgc values are required for transition metal QDuples containing other ligands besides aquo or large organics. This QOuld enable ligand influences on A911: to be discerned. The dearth of data on other ligand systems can be traced to a lack of suitable redox couples. In aqueous solution, most transition metal ion complexes .19 which contain common unidentate ligands are not stable in both oxida- tion states. In at least one of these states, the complex will be unstable with regard to dissociation, chemical oxidation, or other reactions. These circumstances preclude the measurement of the equili- briuxn cell potentials of these couples. Furthermore, the determination of ionic entropies is prevented since these quantities are frequently obtained from the temperature dependences of cell potentials (40). This situation has recently been significantly improved when it was determined that several ruthenium(III)/ (II) complexes were substi- tutionally inert in both oxidation states, at least on the time scale of cyclic voltammetry and other electrochenical perturbation techniques (52-56). Some osmium(III)/(II) amine couples have been found to have 8ll-Inilar properties (57,58). The Ru and the Os couples are electro- c1lenically reversible (i.e., have fast heterogeneous charge-transfer k1|.netics), and as a consequence, cyclic voltammetry can be used to 0'btain accurate values of the equilibrium cell potentials. Thus, the As;c values for these Ru(III)/(II) and Os(III)/(II) couples should be experimentally accessible. Aside from their intrinsic Value these data could also find application as a means of estimating AS°C values for other systems. The reaction entropies for Cr(III)/(II) end Co(III)/ (II) complexes (which cannot be determined directly) might be roughly approximated by the As:c for the analogous Ru or Os system. This procedure is not without some justification. Empirical correla- tions (41-44) suggest that As:c values for simple transition metal c=<>uples are chiefly functions of the type of ligand involved and of the charge on the cation. The actual identity of the central metal. ion appears to be of less significance. Given the paucity of data at 20 present, it is obvious that a generalization such as this cannot be either verified or refuted until further investigations are carried out. C. Homogeneous Electron-Transfer Kinetics of Transition Metal Complexes The great practical and theoretical importance of homogeneous redox reactions has encouraged extensive research in this area. lat , In particu- much attention has been devoted to the kinetics and mechanisms of these reactions. Significant progress has been made for those reac- tions where oxidation-reduction is actually brought about by electron transfer (as opposed to reactions which occur via atom transfer or some, other complex mechanism). The mechanistic simplicity of such e‘-1-ec:tron-transfer reactions has been one of the primary reasons behind the progress in this field (59,60). Probably the most significant and useful advancement in the area of electron-transfer mechanisms is the concept of outer-Sphere and inner- Sphere processes (59). The basic notion here is that an electron- transfer reaction can be placed in one of two broad classifications on the basis of the geometry of the transition state. In an outer-sphere Inechanism, the coordination shells of the oxidant and reductant remain intact while electron transfer occurs. Although some perturbation of the coordination shell is allowed (and usually necessary as will be Seen later), no bond rupture takes place in an outer-sphere process. In contrast, an inner-sphere mechanism involves definite changes in the coordination shells of the reactants. For transition metal com- Plexes, these changes often yield a species in which one ligand is eoznmon to the coordination spheres of both reactants. In this arrange- tllent, the ligand serves to connect the two metal centers. Consequently, 21 the mechanisms of reactions which proceed via such transition states or intermediates are frequently termed "ligand-bridged mechanisms", and one commonly finds that "inner-sphere" and "ligand-bridged" are used synonymou sly. Because the understanding of a redox reaction can be greatly improved by a knowledge of its mechanism, one finds that the charac- terization of a reaction as inner sphere or outer sphere is one of the main concerns of the redox kineticist. The assignment of a mechanism can be straightforward, difficult, or impossible, depending on the ehethical properties of the reactants (59) . When both reactants are a\lbetitutionally. inert on the timescale of the redox reaction, an ontier-sphere mechanism is the only one possible (60). The same res- tit‘iction applies to a reaction containing one inert and one labile I'-’eactant as long as the inert species has no unoccupied binding sites which could coordinate with its labile reaction partner (60) . Thus a cognizance of the ligation kinetics of a transition metal complex is a crucial requirement for an investigation of its electron-transfer I‘eactions. In the case of the present study of the redox kinetics of the Eu cryptates, one sees that a knowledge of the dissociation kinet- ics of these complexes will have both intrinsic worth and a direct bearing on the characterization of the electron-transfer mechanism. A consideration of the substitutional inertness of the reactants is not the only means by which one may distinguish outer-sphere and inner-sphere reactions. One of the first methods employed was the meanination of the products of the redox reaction. Taube's pioneering efforts in this field (61-63) and also in the general area of redox 1(Zinetics and reaction mechanisms must be acknowledged at this point. 22 The product analysis method takes advantage of the substitutional inertness or lability of the reactants and products of certain redox reactions. This can be best illustrated by examining one of these 2+ with Cr2+: reactions; e.g. the reduction of Co(NH3)SCl + Ce(x~ui3)5c12+ + Cr2+ EL Co“ + 01:012+ + 5NH4+ (3) This is a rapid reaction. The Co(NH3)SClz+ and CrC12+ complues are substitutionally inert and only aquate slowly. On the other hand, 2+ Co and Cr2+ are labile species. The preceding points lead one to conclude that this is an inner-sphere reaction which involves the for- mation of a chloride-bridged intermediate or transition state 13: 2+ 2+ ' 4+ (NH3)SCoCl + Cr(OHz)6 -> (NH3)SCo§lCr(OHZ)5 +1120 (4) The lability of the or2+ ion allows the bridged species B to be pro- duced. An intramolecular electron transfer occurs within B to yield Co(II) and Cr(III). This followed by the cleavage of B to yield the reaction products: (NH ) CoClCr(OH )4+ 3+ c 2+ + Cr(OH ) c12+ + 51m+ (5) 3 5 2 5 * ° ' 2 5 4 The lability of the Co(II) ion combined with the substitutional inert- tless of Cr(III) leads- to the rupture of the Co-Cl bond and to the chloride being frozen into the coordination sphere of the Cr(III) ion. F'or the sake of argument, one can imagine that reaction (3) could pro- 2+, Cr3+ and free Qeed via an outer-sphere electron transfer to yield Co 2+ (31-. A subsequent ligand-substitution reaction yields the CrCl QOrnplex which is found in the product analysis: Cr3+ + c1” + c::c12+ (6) 23 Such an outer-sphere route can be ruled out because of the substitu- tional inertness of Cr3+ and rapidity of the redox reaction. The elec tron-transfer reaction would occur and the analysis of the products would be carried out well before reaction (6) could occur to any sig- nif 1cant extent. Thus an inner-sphere mechanism is the only one which is consistent with the properties of the reactants and products. Add itional confirmation was obtained when the reaction was carried out in the presence of 3601-. None of this isotope was found coordinated t0 the Cr3+ ion (59,61-63) which reaffirms the assertion that equation (5) and hence the outer-sphere pathway do not contribute significantly to the production of CrClz+. Because many redox reactions involving Cr2+ are rapid and since Cr3+ is substitutionally labile, several other mechanistic studies have elllployed Cr2+ as a reducing agent for Co(III), Cr(III), and Fe(III) c(unplexes (59,64) . These reactions can be written as Cr2+ + MIII xn+ + Crux Xn+ + M2+ (7) ‘ihere X is the potential bridging ligand. In each case the detection <>f (:an+ has served to verify the presence of an inner-sphere pathway. Reducing agents other than Cr2+ have also been used. Co(CN)§- and 5hu(N83)5 OH:+ are similar to Cr2+ in that they aquate very slowly after <>lxidation. Thus the technique of separation by ion exchange followed 13}? spectral identification. which was used on the products of the Cr2+ ITeactions, can be applied equally well to the products of the Co(CN)§- Eltld Ru(NH3)SOH§+ reactions (59,65,66). Other mechanistic studies have employed Fe2+ or V2+ as reductants. while the product species (Fe3+ or V3+) hydrolyze rapidly, the 24 appropriate remnants of the ligand-bridged intermediates can be observed using flow techniques (59). Another method for identifying an inner-sphere mechanism is to detect the bridged species itself. This is unambiguous evidence only when both of the products are substitutionally inert (59). Thus the reactions of Cr2+ with Ru(III) chloride complexes have yielded I Cr II - Cl - Run species (67,68). Other reactions have been charac— ter :Lzed as inner-sphere by this method (59). One may sometimes diagnose reaction mechanisms from a careful inspection of rate constants or activation parameters. For example, the activation parameters for ligand substitution on the V2+ ion are eJamparable to those found for the reactions between V2+ and certain OXidizing agents. It is strongly suspected that an inner-sphere electron—l:ransfer mechanism is Operative and that the rate-determining Step is the displacement of an aquo ligand on V2+ by the bridging ligand (69,70). Second-order rate laws of the form: Rate - k [reductant] [oxidant] (8) are usually found for both inner-sphere and outer-sphere reactions. Thus no mechanistic assignments can be made when such experimental behavior is observed. However, when dealing with reactions where at least one of the reactants has an aquo ligand, one frequently finds that the rate constant k shows an inverse dependence on [11+]. Often One the rate constant is found to be the sum of two contributions. -1 Of these is an acid-independent term while the other is the [11+] term. This inverse term can be a very significant fraction of the mcperimentally observed rate constant, and in some reactions, the 25 contribution of the [H+] term is so large that the [fin-independent term cannot be reliably determined. The inverse-acid term is generally attributed to the presence of an inner-sphere mechanism which proceeds via a hydroxide bridge (71-73). In more explicit terms, the reaction involves the deprotonation of one of the aquo ligands (this pre- 1 equilibrium step is the source of the [11+] term), followed by the formation of the hydroxide-bridged intermediate or transition state, and finally concludes with the electron-transfer step and any other following chemical process. Therefore, the [Kn-dependence of the rate constant may provide a clue to the mechanism as long as an aquo ligand is present on one of the reactants. While determinations, of mechanism are very important to the study of redox reactions, one is still left with the problem of formulating quantitative explanations for the experimentally observed rate constants. Considerable theoretical work has addressed this question and signifi- cant progress has been made, especially in the case of outer-sphere reactions (59,60,74). The calculation of the energy barrier to electron transfer is simplified for outer-sphere processes because of the absence of bond formation and breakage (59,60). Several theories dealing with this area have emerged (74) including those of Hush (75) and Levich and Dogenadze (76), but the treatment of Marcus has been the most suc- cessful and widely used (77-79). The Marcus theory may be approached on two levels. It can be used in an absolute sense to calculate outer- 8phere electron-transfer rate constants without the use of other rate data, or it can serve as a relative theory which begins with some eatisting rate parameters and employs these to calculate the corres- ponding values for other reactions (60). 26 A summary of the Marcus theory can be started by considering a self-exchange reaction, i.e., one in which the oxidant and reductant are from a single redox couple. kll mums):+ +'m(m3)§+ + Ru(NH3)§+ + Ru(NH3):+ (9) is an example of such a reaction. Obviously no net chemical reaction occurs, and consequently AG° - O for self-exchange reactions. The absolute Marcus theory attempts to obtain the self-exchange rate constant k11 by calculating the associated free energy of activation. In order to do this, the reaction is assumed to be weakly adiabatic. This implies two conditions. First, the interaction between the two reactants is sufficiently small so that neither is directly perturbed by the presence of the other. Thus the activation of the two reaction centers occurs independently. The second condition requires that the interaction between the reactants is strong enough to allow K, the probability for electron transfer in the activated complex, to approach unity. The requirement of weak adiabaticity may seem overly restric- tive, but fortunately most redox reactions which have been studied do indeed satisfy this condition (80). Before continuing on, it should be pointed out that this section is only considering thermal electron-transfer reactions, i.e., those which occur without the absorption of light. Only thermal energy is used in surmounting the activation barrier. Furthermore, the Franck- Condon principle applies in these systems. In more specific terms, because electronic motion occurs much more rapidly than nuclear vibra— ttion, the reactants have no time to change their nuclear geometry while 27 the electron-transfer act takes place. Consequently, there are actually two activated complexes, one for the reactants and the other for the products. These are isoenergetic states which differ only in the loca- tion of the transferred electron. In the reactants' activated complex, this electron is located on the reducing agent, while in the products' activated complex, the electron is found on the reduction product. The laws of thermodynamics require electron transfer between iso- energetic states. In principle, a nonisoenergetic transfer could be employed to circumvent the First Law (74). Having dealt with these preliminaries, one can now consider the free energy of activation. It is helpful to think of this quantity as a sum of contributions which stem from the individual events which make up the activation process. The sequence begins with the diffusion of the reactants toward each other to form the so-called precursor complex. The energy involved here is the electrostatic work (and other forms of work) which must be done when bringing the reactants together from an infinite separation to one which corresponds to the internuclear distance in the activated complex. This term can be quite appreciable, particularly in those cases when like-charged species are being brought together. Once the reactants are in the precursor state, the remainder of the activation process occurs. Changes in the nuclear coordinates must occur in order to reach the activated complex. The isoenergetic con- straints which were described above dictate the nature and the degree of these changes. Both the inner coordination sphere of the reactants and the surrounding solvent must be modified appropriately. The ennergies associated with these processes are respectively designated 28 the inner-shell and the outer-shell terms. The inner-shell term has its origins in the distortions in the metal-ligand bond distances and the changes in ligand conformation which are necessary to attain the activated complex. For example, in the reduction of a trivalent ion to its divalent form, the M(III)—L bond would be required to stretch to a length which was intermediate to those for M(III)-L and M(II)-L. The same considerations apply to the other reactant. The outer-shell term may be regarded along similar lines. The solvent orientation in the activated complex will be intermediate to those surrounding the reactants in the precursor complex and the products in the analogous successor state. One can summarize the preceding discussion by beginning with the relationship between the self-exchange rate constant k11 and the free * energy of activation AG : kll I Z exP(-AG*/RT) (10) where Z is the bimolecular collision frequency, R is the gas constant, and T is the absolute temperature. As discussed previously, the free energy of activation can be expressed as the sum of contributions from each of the steps in the activation process * * in + AGO“It (11) * AG 8 w ‘+ AG r where wk is the work required to form the precursor state from the a * separated reactants, and AGin and AGout are the inner-shell and outer- shell terms described above. It is important to note that the Marcus formulations are associated vwdth a kinetic model in which each collision between reactants of the 29 proper energy and orientation leads to reaction. This differs from the transition-state or ion-pair pre-equilibrium model which is widely applied to solution reactions. The Marcus model is concerned with the non-equilibrium.energy fluctuations of the system rather than a formal equilibrium.between the transition state and the reactants. A recent paper by Brown and Sutin (81) has addressed these two models with regard to the self-exchange rate parameters for some Ru(III)/(II) complexes. The Marcus effective-collision model was found to agree more closely with the experimental results. Analytical expressions for each of the terms in equation (11) have been derived. If the work in forming the precursor state is considered to be purely electrostatic in nature, the appropriate expression for wf can be obtained from Debye-Hfickel theory (81): z z ezN w - ‘ 1 2 1 (12) 583(1+B§u l2) where 21 and 22 are the charges on the reactants, e is the electronic charge, N is Avogadro's number, 88 is the (static) dielectric constant of the solvent, 3 is the distance between the centers of the reactants in the precursor state, B is the Debye-Hfickel parameter (82) and u is ionic strength. The inner-shell term is given by (74,79,81): * 6 f1f2(Aa)2 Acin - 2(f1+f2) (13) where f1 and f2 are the force constants for the metal-ligand bonds in the reactants and As is the difference between the equilibrium M-L bond distances in these two Species. The metal-ligand bonds are assumed to behave harmonically. The expression for the outer-shell 3O term is (74,79,81): * e2 1 1 1 1 1 Acout T (2a; + 232 " 3') (cap ’ F8) (14) where a1 and a2 are the radii of the reactants, sop is the optical dielectric constant (equal to the square of the refractive index of the solvent) and the other terms have already been defined above. In this case, the solvent is assumed to be a dielectric continuum. As formulated here, the absolute Marcus theory is a remarkable achievement since it allows the calculation of a rate constant from non-kinetic quantities. The theory has performed well in qualitative comparisons between related complexes and has proved useful as a con- ceptual structure within which activation barriers may be treated in terms of separate components (60). As a tool for making quantitative predictions, the Marcus theory has met with partial success (60). Reasonable agreements between calculated and experimental self-exchange rate constants have been observed for some systems (81) but this is not always the case (60). One point which is of particular interest to the present work involves the entropy of activation. Many outer- sphere self-exchange reactions have been found to have considerable entropies of activation (AS*‘¥ -30 e.u.) (83). A AS$ value of about -20 e.u. is expected at zero ionic strength but theory predicts an essentially negligible value at the ionic strengths employed for most experimental studies (60). This question will be discussed in Chapter V and a possible explanation will be advanced. While the Marcus theory has encountered problems when used in absolute terms, it has been quite successful when employed in a relative J 31 sense. For the purpose of illustrating the relative theory the following reactions will be considered. k11 3+ 2+ 2+ 3+ Ru(NH3) 6 + Ru(N'H3) 6 + Ru(NHB) 6 + R11(NH3) 6 (15) ' 3+ 2+ 1‘22 2+ 3+ Fe + Fe + Fe + Fe (16) 3+ 2+ 1‘12 2+ 3+ Fe + Ru(NH3)6 +’ Fe + Ru(NHB)6 (17) Reactions (15) and (16) are examples of the previously described self- exchange reaction. Reaction (l7) differs from the other two in that its reactants are from two different redox couples. Such reactions are designated "cross reactions" (59). The self-exchange rate constants can be considered to reflect the intrinsic reactivity of the single couple which is involved. Rate constants for cross reactions are a less clear-cut measure of reactivity since they are composites of the intrinsic reactivities of the oxidant and reductant combined with the effect of the non—zero AG° on the activation barrier. The achievement of the relative Marcus theory has been its ability to relate the cross-reaction rate constants to those for the constituent self-exchange reactions. The activation free energy of the cross reaction, AGtz, can be considered to consist of contributions from the intrinsic (self-exchange) barriers, 0.5Ath and 0.5AG22, combined with 0.5AG:2, the lowering of the activation barrier due to the favorable free energy of reaction, AGiz. The assumption of weak adiabaticity removes the need for any coupled activation of the redox centers. Thus, one finds independent additive contributions from each redox couple involved. In summary, the activation barrier of a cross 32 reaction can be given by (60): * * * o A612 I 0.5(AG11 + AGzz + AG12) (18) This equation can also be expressed in terms of rate constants (59,60,74,79), which yields the Marcus cross-relation: 1. a 2 R12 (kll k22 K12 f), (19) where K12 is the equilibrium constant for the cross reaction and f is given by log f = (103 K12)2 4/ 4 108 (k11k22/22) (20) The f term is a correction factor which accounts for the parabolic shape of the activation barrier. (Equation (18) is a simplified form which does not include this term.) The cross-reaction rate constants calculated using equation (19) have generally agreed quite well with experimental results (59,60). While this is a substantial accomplishment, some difficulties remain. The cross-relation does not work well for reactions involving Co(OHz):+ (60). Detailed studies have isolated the thermodynamic (A632) and intrinsic (Ath and A622) factors which.make up Actz. While the thermodynamic term tends to comply with equation (18), some exceptions have been found for the intrinsic contributions (60). A possible explanation is that nonelectrostatic work terms resulting from the formation of precursor state may not always cancel each other completely (77). The limitations of the relative Marcus theory will be examined in greater detail in Chapter V and some possible sources of these problens will be discussed. CHAPTER II EXPERIMENTAL ASPECTS 33 A. Apparatus Electrochemical measurements were carried out in conventional glass cells having working and reference compartments separated by a "very fine" or "ultrafine" grade frit obtained from Corning Inc. These frits had an average porosity of 1-3 um and prevented significant mixing of the solutions in the two compartments in the 2-3 hours required for most experiments. The capacity of the working compartment was in the 5-10 ml range. The working compartment, the liquid junction formed in the frit, and part of the salt bridge between the working compartment and the reference electrode were surrounded by a jacket through which water from a Braun Melsungen circulating thermostat could be circulated. This enabled the temperature within this region to be controlled to within i0.0S°C. A separate jacket around the reference compartment enabled the temperature of this part of the cell to be controlled inde- pendently when necessary. Bulk electrolyses were performed in cells in which the counter electrode was given its own compartment adjacent to the working compart- ment but separated from the latter by means of a frit. For applica- tions which required substantial current densities, a similar cell with more porous frits was employed to minimize resistance problems. B. Electrochemical Techniques 1. Common Features All solutions were deoxygenated by bubbling with prepurified nitrogen or argon which had been passed through a series of wash bottles containing V(II) solution. The final bottle in the series 34 35 contained water and served to humidify the gas so that prolonged bubbling would not result in extensive evaporation of the solution in the working compartment. Saturated aqueous calomel reference electrodes (SCE) were used throughout this work, although the choice of the fill solution was dic- tated by the experimental conditions. In.most cases, a SCE filled with saturated KCl (KSCE) was used. In concentrated perchlorate media, a KSCE suffers from problems arising from the limited solubility of KClOA. Precipitation of this salt in the liquid junction of the electrode can lead to spurious potentials. Therefore a SCE filled with saturated NaCl (NaSCE) was employed when working with perchlorate solutions. The potentials in this work are quoted versus the KSCE although in a good many cases they were actually measured versus a NaSCE. A platinum wire served as the counter electrode for most experi- ments. When larger currents were passed, a platinum gauze was used. 2. Cyclic VOltammetry Cyclic voltammograms were obtained by means of a PAR l74A polarographic analyzer (Princeton Applied Research Corp.) coupled with a Hewlett-Packard HP 7045A X-Y recorder. Sweep rates in the 50-1000 mV/sec range were used in this work. Under these conditions and with the above instrumentation, peak potentials could be measured with a precision of il-Z mV. Several working electrodes were used in the cyclic voltammetric studies. For redox couples with relatively negative standard poten- tials, a commercial (Metrohm Model E410, Brinkmann Instruments) 36 hanging mercury drop electrode (HMDE) was used. An electrode area of about 0.03 cm2 was employed for most studies. Mercury has a limited anodic range, so for work at positive potentials (E > +200 to 300 mV vs. KSCE), a glassy carbon or platinum "flag" electrode was employed. The latter consisted of a 2 mm2 sheet of platinum foil spot-welded to a fine platinum wire. The Pt flag electrode was pretreated by immer- sion in warm l:l BN03 followed by activation by passage over a Bunsen burner flame. The literature procedure for the pretreatment of glassy carbon electrodes (84) was found to do very little to improve their performance as measured by the residual current and the anodic limit. Therefore, these electrodes were not chemically pretreated before use. 3. Polarggraphy The instrumentation used for cyclic voltammetry was also employed to obtain dc polarograms. Sweep rates between 0.5 and 5 mV/s were utilized. The working electrode was a DME (dropping mercury electrode). Mercury flow rates of l-2 mg/s and column heights of v~50 cm were used. The drop time could be mechanically controlled to be 1, 2, or 5 sec. by means of a PAR 174/70 drOp timer. 4. Preparative Electrolyses Constant-potential electrolyses were performed with a PAR 174A or a PAR 173 potentiostat. The latter was used when currents greater than 10 mA.were required. A stirred Hg pool served as the working electrode while a Pt gauze electrode was used as the counter electrode. The electrolyzed solution was kept under a blanket of deoxygenated nitrogen or argon to prevent any reactions with atmospheric 02. 37 5. Potentiometric Methods A Corning Digital 109 pH meter combined with a regular glass electrode or a miniature combination pH electrode was used for pH measurements and potentiometric titrations. The study of fluoride ion association employed an Orion Model 94-09A fluoride ion-selective electrode. C. Materials 1. Cryptates Solid cryptate samples were obtained from Dr. 0. A. Gansow‘s group. The syntheses of Eu(2.2.l)Cl Eu(2.2.2)C13, Eu(2.2.l)(NO 3' 3’3' Eu(2.2.2)(N03)3, Yb(2.2.l)C13, and Eu(23.2.l)Cl3 involved mixing the appropriate Eu or Yb salt with the cryptand in a nonaqueous solvent such as acetonitrile. After this mixture is refluxed for a few hours, the complex is precipitated from solution by the addition of ethyl ether (85). The Eu(2.2.l), Eu(2.2.2), and Yb(2.2.l) complexes can also be syn- thesized in aqueous solution. One can prepare aqueous Eu2+ and Yb2+ 3+ 3+ by electrolyzing solutions of Eu and Yb in 0-1.§.Et NClO4 (tetra— 4 ethylammonium perchlorate) at -l.l and -l.4 V, respectively. A slight excess of the appropriate cryptand is then added to yield the cryptate complex. This divalent form of the cryptate can then be oxidized to the trivalent state by an electrolysis at the appropriate potential (about 100-200 mV anodic of the formal potential of the cryptate given in Chapter III). The complexes made in this fashion display the same electrochemical behavior as that found for the solid cryptate samples. The yield of Yb(2.2.l)2+ was somewhat low due to the hydrolysis of Yb2+ 38 which occurs in the basic solution produced by the addition of the cryptand. The (2.2.2), (2.2.1), and (2.1.1) cryptands were obtained from PCR, Inc. 2. Reagents and Solvents Except where otherwise noted, the reagents employed in this work were analytical grade and used without further purification. Samples of Et4N0104 (from Eastman or G. F. Smith) used in making sup- porting electrolytes were recrystallized from water. This resulted in substantial reductions in the background polarographic current. water was purified by the use of a Milli-Q purification system (Millipore Corp.). Spectral-grade nonaqueous solvents dried over molecular sieves were used in the nonaqueous studies. Mercury which had been triply distilled under vacuum (Bethlehem Apparatus Co.) was employed in the HMDE and DME and as the pool electrode in electrolyses. 3. Transition Metal Complexes and Compounds Solutions of V2: were synthesized as follows: a solution of V0; is made by dissolving V20s in an appropriate acid (usually HC104). At least a five-fold excess of acid must be used if the dissolution of the V205 is to occur in a reasonable length of time (i.e., less than l-2 days). Electrolysis of the V0: atmosphere at a mercury pool yields Viz. The reduction consumes H+, at -1100 mV under a nitrogen and this must be remembered when a knowledge of the H+ concentration is necessary. The V2: solution produced using this method can be converted to V3: by electrolyzing at -300 mV. The preparation of solutions of U2: involved electrolysis of uranyl perchlorate (obtained from G. F. Smith Co.) in 0.5 M HClO 4 at a stirred mercury pool at ~1000 mV. A subsequent electrolysis 39 at -650 mV yielded ng. A.nitrogen atmOSphere was maintained over the solutions throughout the electrolyses. Solutions of Eng: were obtained by dissolving Eu203 in an appropriate acid. In order to dissolve the oxide within a reasonably short period of time, a stoichiometric excess of acid must be used. The Fe(bpy) §+, Fe(phen) §+, Co(bpy) 3+, and Co(phen) §+ complexes (where bpy - 2,2'-bipyridine and phen - l,lO-phenanthroline) were pre- pared by mixing the metal ion with at least a five-fold excess of the ligand. These complexes were generally studied in a KCl supporting electrolyte since they were only slightly soluble in perchlorate media. Ru(NH3)6-Cl3 (obtained from Matthey BishOp) served as the starting material for the syntheses of several other ruthenium ammine complexes. 3)6-Cl3 in 1:1 HCl for 4-5 ncs-(c104)2 was synthesized using the procedure of Ru(NHB)5Cl°Cl2 was made by refluxing Ru(NH hours. Ru(NH3)5 Lim et a1. (52). A 10-15 m_l_4_ solution of Ru(NH Cl-Cl in 3’5 2 300011 was made. An electrolysis at -700‘mV under 2+ 2 C1. by H20 occurs at an appreciable rate with divalent Ru. The ten— o.1 g NaCF coo, 1m._M_ CF 3 an argon atmosphere yielded the Ru(NH3)50H complex. (Substitution of dency of Ru(II) to form dinitrogen complexes necessitates the use of the argon atmosphere.) A ten-fold excess of NaSCN was added, and the color of the solution changed from faint yellow to red~orange, indicat- ing the formation of Ru(NH3)5NCS+. An electrolysis at -80 mV converted 2+ this divalent complex to Ru(NH NCS The solution was then removed 3’5 from the electrolysis cell and Ru(NH3)5NC8-(C104)2 was precipitated out by the dropwise addition of a slight excess of 5 M NaClO4. The Ru(NH3)SOH3+ complex was prepared by two methods. The procedure 40 of Lim.et al. (52) begins by electrolyzing a solution of 2-3 WE. Ru(NHa)5Cl-Cl in 0.2 _M_ NaCF 2 3 to yield Ru(NH3)SOH§+. This solution was removed from the electrolysis C00, 1 my CFBCOOH under argon at -700 mV cell and added to a solution of AgCFBCOO which had been made by dis— solving Ag20 in CF COOH. The silver solution had sufficient Ag+ to 3 precipitate out all Cl- and to oxidize the divalent Ru complex to Ru(NEB)SOH§+. The resulting solution was filtered to remove AgCl and Ag. The electrolysis cell was cleaned in order to remove any residual 01-, and the filtrate was returned to the cell. An electrolysis at 0.0 V reduced any excess Ag+ in solution. Thus, one should obtain a solution of Ru(NH3)SOH§+ in NaCF COO at the end of this procedure. 3 A.nodification of this method served as the second means for synthe- 3+ 2 the solution was kept in the cell. A second electrol— ysis was then carried out at -50 mV. This oxidized the Ru(NH3)50H§+ produced in the first electrolysis to Ru(NH3)SOH§+, Neither method was 2+ sizing the Ru(N33)SOH complex. After the initial electrolysis of the Ru(NH3)5Cl-C12, ideal since both yielded solutions containing a slight Ru(N33)5C1 contamination. Fortunately, the standard potentials of the 2+/+ 3+/2+ Ru(N83)5Cl and Ru(NHB)SOH2 couples are sufficiently well separated so that the latter complex can be studied without significant 2+ interference from the former. Solutions of the Ru(NH3)SOH complex 3+ 2 O The Ru(NHa)5 py2+ complex (where py - pyridine) was synthesized as were made by adding base (NaOH) to solutions of Ru(NH3)50H follows: a 1 mg solution of Ru(NH3)5Cl-Cl in 0.1 M NaCF C00, 2 3 l mM_CF3COOH was electrolyzed at -700 mV under an argon atmosphere for about 3% hours. The resulting light yellow solution of Ru(NH3)SOH§+ turned a bright yellowbgreen upon the addition of 41 2.5 mMprridine. The color change indicated the formation of 2+ The cis- and trans-Ru(OH2)2(bpy)§+ complexes were made by starting with Ru(bpy)2003-ZHZO (obtained from Dr. G. M. Brown). The carbonate complex was dissolved in l M HpTS (pTS- - p—toulenesulfonate) to yield cis-Ru(032)2(bpy)§+. The high acid concentration is necessitated by the large pKa value of the trivalent form of the cis complex [pKa «a1 based on pKa - 0.85 for the closely similar Ru(bpy)2(py)OH§+ complex (86,87)]. Because it is photosensitive, the cis-Ru(OHZ)2(bpy)§+ complex must be shielded from light. The trans isomer was obtained by exposing the cis form to sunlight. Solutions of the Ru(NHB)4bpy2+ complex were made as follows: a 1.5 my solution of sis-Ru(NH °Cl (obtained from Dr. F. B. Anson) 3)AC12 in 0.1 M NaCF C00, 5 mg! CF3COOH was electrolyzed under argon at ~550 mV 3 for about two hours to yield Ru(NH3)4(OH2)§+. The solution was then made 14 mg in bpy. About 30 minutes are required for the Ru(NH3)4bpy2+ complex to form quantitatively. Several complexes were obtained from outside sources. Ru(NH3)2(bpy)2'(0104)2, dis-Ru(NH3)4(H20)2°(CF3SO3)2, Ru(en)3-Br3, and Ru(bpy)2C03-ZHZO were kindly provided by Dr. Gilbert Brown. Professor Larry Bennett's group supplied a sample of Ru(NH3)4phen-(CFBSO3)2°3H20. Co(sep)Cl3 (where sep - sepulchrate, see reference 88 for further details) was obtained from Professor John Endicott. Ru(bpy)3°C12°6H20 was purchased from.G. F. Smith. The procedure of Bottomley and Tong (89) was used to synthesize Os(NH3)6:I3. A solution of 0.5 g of Na 08016 (from Matthey BishOp Inc.) 2 in 15 ml of concentrated ammonia was made and 0.12 g of NHACl was added 42 to duplicate the solution conditions of the Bottomley preparation. After the addition of 0.5 g of Zn dust, the solution was refluxed for five hours under an argon atmOSphere. (The reaction is 2- 2 208016 + 12NH + Zn +20301113):+ + Zn +I+ 12Cl-). The solution was 3 allowed to cool to room temperature and 3 ml of ammonia were then added. The solid Zn and Os which had been produced were filtered off. The filtrate was cooled in an ice bath, and 2 g of KI were added. The yellow precipitate of Os(NHB)6°I3 was isolated by filtration. The solution cannot be cooled excessively since KI will precipitate along with the Os complex. The synthesis of T10104 involved combining stoichiometric quantities of T12C03 (Alfa Ventron) and HClOa. A solid brown impurity remained after the T12C03 and the filtrate was concentrated to about a quarter of its original had dissolved. This solid was removed by filtration, volume. The solution was cooled and TlC104 was allowed to crystallize from solution. The yield is about 80% of the theoretical value. D. Kinetic Methods l. Cryptate Aquation Studies Aquation rate constants for the Eu and Yb cryptates were obtained by recording polarographic limiting currents or cyclic voltammetric peak currents as a function of time. This yields concen- tration vs. time.data since these two currents (after correction for background) are proportional to cryptate concentration. The timescale of these electrochemical techniques is sufficiently short relative to those of the aquation reactions so that each recorded current is essentially a measure of the instantaneous concentration. The 43 polarographic method has slightly better precision since reproducible electrode areas are more easily attained with a DME than with a HDME. 2. Homogeneous Redox Reactions - Principles Polarography was used in a similar fashion to follow the homogeneous redox reactions of the cryptates. In this case, the measurement is complicated by the presence of the coreactant, which is also an electroactive species. In fact the limiting current cannot serve as a practical measure of the cryptate concentration when both the cryptate and its coreactant are electrochemically reversible. Uhder these conditions, the current recorded at a given potential would contain contributions from both of the redox couples present. For example, difficulties would be encountered if one wanted to follow the progress of a reaction between Ru(NH3)g+ and Eu(2.2.1)2+ by monitoring the polarographic oxidation current of Eu(2.2.1)2+. Depending on the potential region employed for the measurement, the observed current will be either the difference between the Ru(NH3)2+ reduction current and the Eu(2.2.l)2+ oxidation current or the sum.of the oxidative currents of Eu(2.2.l)2+ and the reaction product, Ru(NH3)§+. Similar considerations would apply if one attempted to watch the formation of the Eu(2.2.l)3+ reaction product. In any case, one cannot obtain unambiguous determinations of the concentrations of the reactants or products. ‘ Fortunately, a number of well-characterized inorganic redox couples are electrochemically irreversible since they have, or can be made to have, slow heterogeneous charge-transfer kinetics. If the 44 degree of irreversibility is sufficiently large, one will find a potential region within which neither the oxidized form of the couple is reduced nor the reduced form is oxidized. The appropriate form of this couple can then serve as a reductant or oxidant for a cryptate since the concentration of this last species can be monitored in the potential region described above without any interference from the other couple. Polarography appears to be the best method available for the study of the homogeneous redox reactions in this work. The cryptates and a number of the other coreactants do not have any absorption bands in the visible or near ultraviolet regions of the spectrum which would be suitable for spectrophotometric use. The nature of cyclic volt- ammetry is incompatible to the present situation where both forms of a redox couple are found in solution during the course of the redox reaction. In cyclic voltammetry, one can calculate a concentration from the appropriate peak current; i.e., a potential sweep is required to obtain the desired information. Obviously, a potential scan for this purpose will require the selection of an initial potential which is well anodic or cathodic of the standard potential of the redox couple under study. If, for example, an anodic initial potential is imposed on the system, all of the reduced form of the couple in the vicinity of the working electrode will be converted to the oxidized form. The ensuing sweep will reflect this concentration distribution and not the one existing in bulk solution. Analogous considerations hold if a cathodic initial potential is chosen. The superiority of Polarography stems from the fact that each current measurement is made 45 under essentially constant-potential conditions. The perturbation applied to the system is the growth of the drOp on the DME. In con— trast, cyclic voltammetry perturbs a system by applying a linear ramp of potential. 3. Homogeneous Redox Reactions - Experimental Specifics Five homogeneous electron-transfer reactiens involving Eu + cryptates were studied: Eu(2.2.l)3+ + Euiq, Eu(2.2.l)3+ + Viz, Eu(2.2.l)2+ + Co(NH3):+, Eu(2.2.2)3+ + Ruiz, and Eu(2.2.2)3+ + viz. All kinetic runs were made at 25.0 i 0.1°C. Efforts were made to minimize the amount of cryptate sample used in each run in order to conserve the rather limited supply. In order to facilitate data analysis, pseudo first-order conditions were employed with the non- cryptate reactant present in at least a ten-fold excess. Experiments involving Euiz began with an electrolysis of 1-5 mg Eng: in 20 mg EtéNClOA (u - .l or .2) at -l.2 v to yield a solution of the divalent cation. The adsorption of the Et N+ cation 4 gives rise to kinetic double-layer effects which slow the charge- transfer rates of Eng: and Eu2+. Polarograms of the Eu3+ and Eu2+ aq 89 39 solutions show no faradaic current between -200 and -800 mV. The concentration of the cryptates can be monitored by recording the polarographic current at some potential in this region. After the electrolysis, the homogeneous reaction was initiated by adding a known quantity of solid Eu(2.2.l)(NO3)3 or Eu(2.2.2)Cl3 to the Bug: solution. The latter was stirred during this process and kept under a N2 or Ar blanket. The solid cryptate dissolved in at most 1-2 seconds. This mixing time is not significant relative to the half lives of the 46 redox reactions. The reduction of Eu(2.2.l)3+ and Eu(2.2.2)3+ was followed by recording the polarographic current at -700 mV as a function of time. A correction for the background (charging) current was applied to the recorded data. The potentials used for these measurements were in the diffusion-limited regions of the polaro- graphic waves of Eu(2.2.l)3+ and Eu(2.2.2)3+. From the Ilkovic equation (see Chapter III), the corrected current is directly pro- portional to the cryptate concentration. The acid dependence of the 3+/ 2+ rate constant was studied for both the Eu(2.2.l) and the 3+/ 2+ Eu(2.2.2) reactions. A.similar procedure was employed in the Vi: reactions. A solution of 2-5‘mMLVO; was electrolyzed at -l.2 V to yield Viz. The solution was then made 20 mM in EtaNC104. The Et4N+ ion could not be present initially because the resultant double-layer effects would slow the electroreduction rates of V0; excessively. Ionic strengths of 0.1 were used in the. Eu(2.2.l)3+ - v: and Fu(2.2.2)3"' - vi: reactions, although the former was carried out in a KPF6/HCl/Et4NClO4 electrolyte and the latter reaction used a LiClOl‘lHClOAIEtANClO4 electrolyte. It was not possible to obtain reliable kinetic data for the Eu(2.2.l)3+ - Vi: reaction in concentrated perchlorate media. The homogeneous redox reaction was initiated by adding a known quantity of Eu(2.2.l)(N03)3 V::. The loss of the trivalent cryptate was monitored by recording or Eu(2.2.2)Cl3 (with stirring) to the solution of the polarographic limiting current at about -6OO mV. Significantly more cathodic potentials could not be used since appreciable currents due to V3: reduction were observed in this region. 47 The preparation for the Co(NH3)2+ - Eu(2.2.l)2+ reaction began with the electrolysis of a ~l my solution of Eu(2.2.l)(NO3)3 at -900 mV to obtain Eu(2.2.1)2+. A (C3H7)4NC104/LiC104/HC104 electro- lyte (u - 0.05) was used because the double-layer effects arising from the strong adsorption of the (03H7)4N+ ion slowed the electro— reduction of Co(NH3)z+ to a point where it would not interfere with the observations of Eu(2.2.l)2+. The double-layer effects resulting from Et4N+ adsorption were not sufficient for this purpose. The electrolysis of the cryptate had to be carried out quickly in order to minimize the loss of Eu(2.2.1)2+ which occurs through the dis- sociation reaction: Eu(2.2.1)2+'+ Bug: + (2.2.1). Fortunately, the reduction could be completed in 20-30 minutes. While the electrolysis was being carried out, a 20 my solution of Co(NH °(C104)3 was 3)6 deoxygenated by bubbling with nitrogen. A known volume of this solution was transferred to the Eu(2.2.l)2+ solution by means of a gas-tight syringe. The ensuing reaction was followed by recording the Eu(2.2.l)2+ polarographic limiting current at -300 mV as a function of time. More anodic potentials could not be used because the Eu:: which was produced from the Eu(2.2.l)2+ dissociation would also yield an oxidation current in this region. The low solubility of Co(NH3):+ in perchlorate media coupled with the need to maintain pseudo first-order conditions limited the Co(NH3):+ concentration to the 5-10 m_M_ range. The cryptate redox rate constants and those for cryptate aquation were generally found to have relative precisions of 5-102 as deter- uuned from the average deviation of 3-10 kinetic runs. CHAPTER III ELECTROCHEMICAL STUDIES OF EUROPIUM AND YTTERBIUM CRXPTATE FORMATION IN AQUEOUS SOLUTION 48 A. Introduction The study of the eurOpium and ytterbium cryptates has several interesting features. The trivalent Eu and Yb complexes are the first trivalent cryptates to be scrutinized in aqueous solution (with the 3+’- macrotricyclic cryptate (90)). The comparison exception of a La of the complexation thermodynamics and kinetics of these trivalent cryptates with the corresponding quantities for monovalent and divalent cryptates should provide useful insights into the factors (such as ionic charge and size) which influence cryptate formation. If the electrochemistry of the Eu and Yb cryptates is anything like that for the aquo ions, these complexes will be the first electro- active cryptates to be studied in aqueous solution, and electro- chemical techniques should prove to be quite effective for the investigation of their properties. These methods can be employed to determine the relative stabilities of the trivalent and divalent forms of the Eu and Yb cryptates and to analyze these stabilities in terms of enthalpic and entropic components. In addition, the electrochemi- cal behavior of these complexes will yield some general information on the effect of cryptate formation on the electrochemical reactivity of europium and ytterbium, B. Results 1. Complexation Thermodynamics Cathodic-anodic cyclic voltammograms were obtained for each of the trivalent cryptates. Substantial changes in the electrochemi- 3+/ 2+ 3+/ 2+ cal behavior of the Eu and Yb couples are found when these 49 50 ions are encapsulated within cryptate cavities. The cyclic voltam- mograms shown in Figure 2 are typical. Figure 2a contains the voltammetric waves observed for the reduction and subsequent oxi- dative regeneration of Euzz in acidified 0.5 M NaClO . The results are independent of acid concentration for pH 1-4. The separation between the cathodic and anodic peak potentials is large and sweep- rate dependent, reflecting the electrochemical irreversibility (i.e., the slow heterogeneous charge-transfer kinetics) of the 3+] 2+ aq mograms for Eu(2.2.1)3+ in 0.5‘1\_‘I.NaC104 shown in Figure 2b have a much Eu couple (91). On the other hand, the corresponding voltamr smaller, sweep-rate independent peak separation of 65 mV, which is quite close to the 57~mV value expected for electrochemically revers- ible, one-electron couples at room temperature (92). Similar peak separations are found for the other cryptate couples studied. The observed peak separations demonstrate that the heterogeneous electron-transfer rate constants of the Eu cryptates are at least two orders of magnitude larger than those for Buzz/2+. In order to yield reversible cyclic voltammograms with the sweep rates used in these studies, the Eu cryptates must have kS values of at least »~.01 cm.s-1, where k8 is the electrochemical charge-transfer rate constant at the standard potential of the couple of interest (92,93). 3+/ 2+ In contrast, Eu aq 5 has been found to have a ks value of ~8 X 10- cms-1 (94). Thus a considerable alteration in the electrochemical reactivity of the europium couple occurs when it is encapsulated within a cryptate cavity. 3+/ 2+ 3+/ 2+ The formal potentials E for Eu(2.2.1) and Eu(2.2.2) f 51 Figure 2. Cathodic-anodic cyclic voltammograms of so: and Eu(2.2.l)3+. Key: (a) 0.4 mM Eu3+ in 0.5 M NaClO4 at pH 2. (b) 0.35 m! Eu(2.2.1)3+ in 0.5 M NaClOa at pH 7. Electrode area - 0.032 cm3. Sweep rates: (1) 200 mV 5-1; (2) 50 mV s-l. 52 I . T r l l T I l o) Eu3+(oq)/Eu2+(oq) (I) 2*- .. (2) | _ — O- _ -l ..... _. i(/.LA) 2 b) Eu(2.2. I)3+/Eu(2.2.|)2+ _‘ (I) '” (2) - I 0 " '1 -l >— _ l I 1 I L l l L 200 400 soo 800 900 -E (mv. vs. S.C.E.) Figure 2. 53 obtained from the mean of the cathodic and anodic peak potentials (95) were found to be markedly more positive than for Fuzz/2+. Analogous 3+/2+ (96) and Yb(2.2.l)3+/2+ behavior was also observed for Eu(23.2.l) redox couples. These results are summarized in Table 1, along with data for the Eu and Yb aquo couples. Identical cyclic voltammograms with peak separations of about 60 mV were obtained for the cryptate couples in the absence and presence of added cryptand over the range of sweep rates 50-500 mV 8-1, for solutions in the pH 1-9 range, and for supporting electrolytes containing sodium, barium, or tetra- ethylammonium cations which have widely differing abilities to compete for the cryptand (6). These results indicate that both the trivalent and divalent cryptates are substitutionally inert on the cyclic voltammetric timescale. Since the difference in formal potentials E f between the complexed and uncomplexed one-electron redox couples can be related directly to the difference in the free energy driving force for these processes, we can write AE - (2.303 RT/F) log(KII/KIII) - (AG f AGEI)/F (1) o - III where AG° and AG° are the free energies of complexation for the III II trivalent and divalent species, respectively, and KIII and KII are the corresponding stability constants. The resulting values of (AG are listed in Table l. O 0 III ' AGII) It is seen that the stabilities of the trivalent cryptates are consistently and substantially smaller than those of the corres- ponding divalent cryptates; i.e., the values °f (A6111 0 AGII) are 54 c mm case oeooz.mw.o e e .. O O u 0 O O me o as e a nose oeoz m ea c +N\+mae N Moo» 0 u. m n— o as muse in m V was 2H o +~\+mow s .. m 0 m 0 O O D o m ohm sec 2 2m o +~\+mAH N No m Hm mHN ooaooz.mn.o o a *V q ll. u 0 O O : e~ a he a s new oeoz m ea c +N\+mA~ N No a m.e~ one is-“ mevooeooz.mw.o mNo ooeooz m:.o . . e o I. . u . . . a m cm a co o s «we oeoz m as o +N\+mAH N No m I. oo we ems Am mesmeeoz ze.o +~\+m=m .:.o Hhaoa.Hoox HhHoE.Hoox .=.o >6 ouhaouuooam oHoaoo Nooom HHH HH HHH HH HHH on Q Om8) where the cryptand is significantly deprotonated. Table 3 summarizes rate constants and activation parameters for the dissociation of five europium and ytterbium cryptates. The progress of the dissociation reactions was followed by periodically monitoring the cathodic or anodic voltammetric peak heights for the trivalent and divalent cryptates, respectively. Concentration-time data for the dissociation reactions were obtained from these cyclic voltammetric peak currents by using (92b): ip - .4463 (nF)3/2 A (Dv/RT)J§ c (5) where i.p is the peak current (corrected for background) in uA, A is the electrode area in cm2, D is the diffusion coefficient of the reacting species in cmzs-l, c is the bulk concentration in mg, v is the scan rate in V-s-1 and the other terms have their usual meaning. Concentrations were also calculated from polarographic 64 Table 3. Rate Constants and Activation Parameters for Dissociation of Europium and Ytterbium Cryptates at 25°C. kda AHIb AS#C -1 d -1 d Cryptate Medium sec. kca1.mol. e.u. 3+ -7 Eu(2.2.1) 0.5g Na0104(pH ~7) 4.1X10 - - 0.5g NaClO4(pH 2.5) 3.0xlo'7 18.9d -25.5d 0.1g Et4NC104(pH 2.5) 4.0xlo'7 19.2d -23.5d Eu(2.2.1)2+ 0.1g Et4Nc1o4(pH 2.5) 2.0xlo’4 15.1e -25.oe Eu(2.2.2)3"' 0.1g Et4NC104(pH ~7) 1.1x1o’3 13.8e -26.oe Eu(2.2.2)2+ 0.055 Et4NC104 3.0xlo’5 18.3‘1 -16.0(1 + 0.033_M_Ba(NO3)2 Yb(2.2.l)3+ 0.5g NaClO 1.3xlo'6 22.3d -1o.5d 4 aFirst-order rate constant for cryptate dissociation at 25°C, obtained at pH values where the acid-independent pathway dominates, i.e., where k& = kd (eqn(9)). bEnthalpy of activation for cryptate dissociation, determined from eqn(7). cEntropy of activation for cryptate dissociation, determined from eqn(8). dDetermined over the temperature range 25-60°C. eDetermined over the temperature range 5-35°C. 65 limiting currents by using the Ilkovic equation (92b): 1 2 1 id=706nDl2 m/3 td/6 c (6) where id is the (baseline-corrected) limiting current in uA, m is the mercury flow rate through the DME in mg 3.1, t is the drop time in d seconds, and the other terms have the same meaning and units as in equation (5). The dissociation reactions of the trivalent and divalent cryptates adhere to first-order kinetics. Plots of In c (or ln ip or 1n id) vs. time are linear over several half lives. The enthalpy and entrOpy i i of activation (AH and AS respectively) were calculated from the temperature dependence of the dissociation rate constant k.cl (101): Mi .. -Rii‘fl‘i - n m d(1/T) AS? - AH*/T + R ln(kBT/h) - R lnflkd (8) where kB is the Boltzmann constant, and h is Planck's constant. (Strictly speaking, k.d is the acid-independent rate constant defined in equation ( 9)) . Experimental conditions were arranged so that the dissociation reactions went essentially to completion. Initially, strongly acidic solutions were employed to achieve this end. However, for some reactions it was found that the apparent first-order rate constants for dissociation ké increased significantly as the pH was lowered. This pH dependence was determined to be in accordance with the rate law: k5 = Rd + half] (9) 66 Similar behavior has been noted previously for the dissociation kinet- ics of univalent and divalent cryptates in aqueous media (27). The derived values of k.d and kH for the dissociation kinetics of four eurOpium cryptates are given in Table 4. The values of kd given in Table 3 were obtained under conditions where the acid- dependent term in equation (9) is unimportant, i.e., where k.d “’ké. For some reactions, this condition could only be achieved by working in neutral media where the cryptand released by the dissociation reaction could prevent aquation of the remaining cryptate. In such cases the occurrence of such back-(complexation) reactions was prevented by adding a small stoichiometric excess of a metal cation such as Ba2+ that will strongly complex the released cryptand (6). Unlike protons, these complexing agents were not found to have any significant effect upon ké. Table 4. Acid Acceleration of EurOpium Cryptate Dissociation at 25°C. a b kd 1% kmfl‘d Cryptate Medium sec-1 M-l sec.1 5.1 3+ £7 . -6 Eu(2.2.l) 1g LiClO4/HC104 3 x 10 ~1.o x 10 ~3 Eu(2.2.1)2+ 1g LiC104/HC104 1 x 10‘4 2.5 x 10"3 40 Eu(2.2.2)3"' 0.1g LiC104/HC104 l x 10"3 0.2 200 Eu(2.2.2)2+ 0.1gLic104/Hc104 -5 x 10'5 7.5 x lo‘3 ..150 aFirst-order rate constant for acid-independent aquation pathway, obtained from intercept of plot of k5 versus [H+] (eqn (9)). b Second-order rate constant, determined from slope of plot of ké versus [n+1 (eqn (9)). 67 The presence of significant concentrations of OH- and F- were also found to have a marked accelerating effect upon the dissociation rates of the trivalent lanthanide cryptates. For example, the addition of 50 mg OH- or F- to 0.5-MNaC104 reduced T55 for the aquation of Eu(2.2.l)3+ at 25°C from 27 days to 22 minutes and 2.8 days, respectively. Table 5 summarizes these effects. Table 5. Anion Effects on Eu(2.2.1)3+ Dissociation. Anion Anion Concentration kd a 09 (sail) on‘ .02 1.6 x 10" .046 5.2 x 10-4 .10 1.3 x 10"3 F‘ .01 1.0 x 10‘6 .05 2.9 x 10‘6 .50 1.4 x 10‘5 aDissociation rate constant for Eu(2.2.1)3fi determined at 25 C. Ionic strength - 0.5. C. Discussion 1. The Effects of Varying the Cation Charge upon Cryptate Thermodynamics In view of the substantial differences in the complexation thermodynamics between corresponding divalent and trivalent lanthanide cryptates (Table 2), it is of interest to compare these results with the thermodynamic parameters for other divalent cations of comparable size, as well as for univalent cations. 68 Table 6 summarizes values of the free energies (A63), enthalpies (AHz), and entropies (A52) of complexation for alkali and alkaline earth cations with (2.2.1) and (2.2.2) cryptands as well as for the lanthanides considered here.and the post-transition metal cation Pb2+. 3+ cryptates were obtained from The values AG; given for Eu2+ and Eu the measured values of KII for the divalent complexes together with the corresponding values of (AGIII - AGII) given in Table 1. Also listed in Table 6 are estimates of the ionic radius of each cation. These values were obtained using the method of Goldschmidt but are from the more recent compilation of Shannon and Prewitt (11) and are estimated for a coordination number of six. Although there are a number of scales of ionic radii which differ to a greater or lesser degree (102a), there is reason to believe that the scale used here provides a reasonable approximation (i0.lA) to at least the relative radii of the bare ions in aqueous solution (102b) as well as the effective cation radii within the cryptate cavities. The use of larger coordination numbers such as eight, which may be appropriate f0r the larger cations considered here, increases the absolute values of the ionic radii by ca. 0.1-0.15 A, but the differences between the values remain relatively unchanged provided that the coordination numbers do not greatly depend upon the cation size (11). The large variations in AG; and particularly in the enthalpic and entrapic components AH; and As; seen between different cations of Group I and II have been discussed by Lehn and co-workers (6,12). The markedly increasing values of -AGZ seen as the ionic radii of the 69 waomv sao.suv oo.s mam.mmv cao.e-v oH.m nn.o +msm mime she.meo om.sa mimic cam.rv os.ma NH.H +Nsm we.oe am.ma he.oa wH.H +~oe no.3- oa.ea o.eo.NH oe.e om.o o.oo.w om.H +~on oN om.oH o.co.oH oe.me oH.o o.oo.oH oH.H +~nm om.oe o~.o o.oo.o oNN oo.~ o.om.s oo.H +~oo was- w~.m so.~ o~.e +wu om.seu om.HH o.oo.n om.o- os.m o.om.m no.H +ce use- oo.HH o~.e oe.s- ow.o o.os.m mm.a +e ox- os.a o.om.m o~.o os.n om.“ No.H +oz we ss.H w~.e os.HH oo.o o.oo.m om.o +en .=.o HhHoE.Hmox deoa.amox .=.o “Hoa.Hcox hHoE.Hmox M cowumu wmc was- was: wmo was- was- on osoocsso A~.~.NV ocmuozuo Aa.~.~v .oomu an consensus Au.~.~v and Aa.~.~o mafia maoauoo Hmuoz maowum> we coauoxwfioaou mo moaoouucm one .moaoncucm .moamuocm monk .o marsh 70 Notes to Table 6. annic radius for metal cation, taken from the compilation in ref. 11 for coordination number six. bFrom ref. 6; determined at ionic strength u - 0.5. cFromref. 12; u ~.0.05. dDetermined from experimental value of KII (see text). eCalculated from A62 for the corresponding Eu(II) cryptate combined with the value of (1101"II - Ash) given in Table 1; u - 0.1. fEstimated value; obtained by assuming A8; to be equal to As; for the corresponding Sr cryptate (see text). 8Calculated by combining As; for the:corresponding Eu(II) cryptate ‘with the value of (ASIII - ASII) given in Table 1; mar 0.5. hCalculated from the corresponding AG; and As; values. iFrom M. H. Abraham, A. F. Danil de Namor, and W. H. Lee, extended abstracts, 29th Meeting, I.S.E., Budapest, Hungary, 1978. 3From ref. 17. 71 cations approach the radius of the cryptate cavity (ca. 1.1 and 1.4 A for (2.2.1) and (2.2.2) cryptates, respectively) could be due in part to steric factors associated with the fit of the cation into the relatively inflexible cryptate cavity (6). However, these variations in AG; consist of marked alterations in As; as well as ABE. This indicates that an important factor determining the sensitivity of cryptate stability to cation size is the partial desolvation of the cation that is required in order for cryptate formation to be accom- plished (12). Thus the large increases in -AH; and -As; that generally occur as the cation radius increases for a given cationic charge (Table 6) undoubtedly stem in part from the decreasingly negative hydration enthalpy and the decreasingly negative hydration entropy of the aquated cations under these circumstances (12). Such influ- ences also appear to be paramount in determining the dependence of cryptate thermodynamics upon the cation charge. Thus cations such.as 2+, or K+ and Ba2 Na+ and Ca + have similar ionic radii, and yet the divalent cations exhibit markedly less favorable values of -AH; for formation of either (2.2.1) or (2.2.2) cryptates. These differences are nevertheless counteracted by markedly more favorable values of As; so that -AG; is typically larger for the divalent compared with the univalent cations. The variations in AH; are in harmony with the markedly more negative hydration enthalpies for divalent cations compared with those for univalent cations of similar size (12,103). Although increasing the cationic charge will result in increasingly favorable ion-dipole interactions with the cryptate ether moieties, 72 these interactions are expected to be relatively less favorable compared with those involving the solvent molecules on account of the smaller dipole moment of dimethyl ether versus water (12). These factors therefore result in a net increase in AH; as the cationic charge increases. The more positive values of As; for the divalent cryptates undoubtedly result from the more negative hydration entrOpies of the divalent versus univalent cations. These stem from a greater degree of "solvent ordering" (or less "solvent disordering") in the vicinity of the former cations (103) which is markedly dimin- ished or removed upon cryptate formation (12). 3+ The differences in both cation size and charge between Eu and Eu2+ are therefore both expected to play important roles in determin- ing the relative stabilities of their (2.2.1) and (2.2.2) cryptates. Thus Eu3+ has a noticeably smaller ionic radius (0.95 A) than Eu2+ (1.17 A) which, along with the increase in the cationic charge, contributes towards a markedly more negative hydration enthalpy for the former cation (103). These two factors taken together appear to account nicely for the relatively more positive values of AH; and As; seen for Eu(III) versus Eu(II) cryptate formation (Table 1). The larger value of (AHIII - AHII) seen for Eu(2.2.2)3+/2+ 3+/ 2+ (17.7 kcal. nol’l) compared with Eu(2.2.1) (10.7 kcal. nol'l) could be anticipated since the AH; values of (2.2.2) cryptates appear to be more sensitive to cation size than those of (2.2.1) cryptates, 3+ at least for cations of comparable size to Eu and Eu2+ (Table 6). For example, AH; for Ca(2.2.2)2+ is 10.2 kcal. mol"1 less negative 73 than for Sr(2.2.2)2+, whereas AH; for Ca(2.2.1)2+ is ca. 4 kcal mol-1 less negative than for Sr(2.2.1)2+. One reason for the behavioral difference between (2.2.2) and (2.2.1) cryptates may be the greater size mismatch between these cations and the cavity of the former ligand. The similar effects seen for the complexation thermodynamics of 3+/ 2+ 3+/ 2+ the other lanthanide couples Eu(23.2.l) and Yb(2.2.l) (Table 1) also support these arguments. The relative instability of Yb(2.2.1)3+ versus Yb(2.2.1)2+, arising again.from a marked enthalpic destabilization outweighing a smaller entrapic stabilization, is expected in view of the smaller ionic radius of Yb3+ (0.86 A (11,18)) compared to r152“ (0.93 A(lS)). Table 6 also lists the absolute values of AG; for Eu(2.2.l)2+ and Eu(2.2.2)2+ obtained from the experimental values of KII’ along with AG; for the corresponding Eu(III) cryptates calculated by combining the divalent results with the data in Table 1. Although the ionic 2+ radii of Sr and Eu2+ are very similar, it is seen that the values of -AG; for the latter cation are significantly larger with both (2.2.1) and (2.2.2) cryptands (by 2.7 and 3.4 kcal mol-l, respectively). These differences may arise from weak covalent bonding between the oxygen and nitrogen atoms of the cryptand with the unfilled 4f-orbitals 2+ on Eu Although the chemistry of divalent lanthanide ions closely resembles that for the corresponding alkaline earth cations having 2+ similar ionic radii (38), the stability constant of the Eu -EDTA 2 complex is also somewhat larger than for the Sr +-EDTA.complex (104). 74 It is interesting to note that Pb2+, which has an ionic radius close to those for Sr2+ and Eu2+ (Table 6), exhibits a value of -A62 for (2.2.2) cryptate formation which is markedly larger than for Sr2+ (Table 6). This difference can also be attributed to the influence of covalent bonding. Although absolute values of AH; and A8; were not determined for the lanthanide cryptates, rough estimates are included in Table 6 for comparison purposes. The entropy values were obtained by assuming that As; for Eu(2.2.l)2+ and Eu(2.2.2)2+ are approximately equal to those for Sr(2.2.1)2+ and Sr(2.2.2)2+. values of AH; were then com- puted by combining these estimates of As; with the experimental values of A02. Although this procedure may be questionable, it is supported by the similar activation entropies observed for the dissociation of 2+ 2+ corresponding Sr and Eu cryptates (vids infra). 2. Association of Lanthanide Cryptates with Small Anions It has been asserted that cryptates provide an unusually effective way of shielding the encapsulated ions from their environ- ment. However, the strong ion association found between the tri- positive cryptates and F- and OH' is a striking example of the limitations of this argument. Particularly unexpected is the finding that both the first and second cumulative association constants 81 and 82 are comparable to or even larger than the values for the corresponding aquo lanthanide cations (Table 2). Therefore it is possible that one or two anions are able to approach closely or even contact Eu3+ or Yb3+ despite the encapsulation of the cations within 75 the cryptate cavity. Inspection of molecular models reveals that such small anions could indeed fit between the polyether strands of the cryptate, although a significant distortion of the symmetrical endb-endb form is required in order to bring two anions into contact with the encapsulated cation. Bearing in mind that the measured association constants reflect the ability of the anion to compete with water molecules for sites adjacent to the metal cation, it is plausi- ble that the surprising stability of the cryptate-anion complexes arises simply from the markedly weaker salvation of the cryptate cation compared with the aquated metal cation. An alternative way of rationalizing the experimental results is to regard the cryptand which surrounds the cation as a region of saturated dielectric and therefore of low dielectric constant which will enhance the coulombic attraction between the encapsulated cation and the incoming anions. It is not clear if the cation-anion associates involve direct cation-anion contact or solvent-separated ion pairs. It therefore may not be necessary for the complexing anions to be buried between the cryptate polyether strands in order to achieve strongly favorable interactions with the metal ion. 3. The Effects of Varyinggthe Cation Charge upon Cryptate Substitution Kinetics The kinetic parameters for the dissociation of lanthanide cryptates summarized in Table 3 exhibit some surprising features. Most prominently, the rate constant k.d for dissociation of Eu(2.2.l)3+ is almost three orders of magnitude smaller than for the corresponding 76 divalent cryptate Eu(2.2.l)2+ despite the substantially lower thermo- dynamic stability of the former complex (Table 1). Even the enormously larger driving force for the dissociation of Eu(2.2.2)3+ compared with Eu(2.2.l)2+[(AG£II - AGII) - 9.7 kcal. mol.-1, Table 1] results in values of ka that are only ca. 30-fold larger for the former cryptate. As in the case of the corresponding thermodynamic parameters, some insight into the factors responsible for this behavior can be obtained by comparing the rate constants and activation parameters for Eu(III) and Eu(II) cryptate dissociation and formation with corresponding data for cryptates of alkali and alkaline earth cations having similar ionic radii. Tables 7 and 8 contain such data for (2.2.1) and (2.2.2) cryptates, respectively. The values of k.c for the europium cryptates were obtained by combining the values of k d values of AG; listed in Table 1. Approximate estimates of the corres- given in Table 3 with the ponding activation parameters AH: and A8: are listed in Tables7'and 8; these were similarly obtained using the estimates of AH; and As; given in Table 6. Inspection of Tables 7 and 8 reveals the importance of the charge as well as the size of the cation in determining the kinetics of cryptate substitution. Thus for the series of cations Na+, Ca2+, Eu3+ which have comparable radii, both k.d and kt fall dramatically as the cation charge increases. For example, Eu(2.2.l)3+ exhibits values of kc and ka that are ca. 107-fold smaller than for Na(2.2.l)+ and ca. 104-fold smaller than for Ca(2.2.1)2+ (Table 7). Since Eu Srz+ 2+ and have very similar ionic radii and also exhibit comparable kinetic parameters with both (2.2.1) and (2.2.2) cryptands (Tables 7 and 8),it 77 s OH- m «N so-onm H mm o +mo» HoH V HH.H~V amm.o m.m- ~.oH sh-onH.s mo.o +msm cl 0 II o o o D HNH V Hm s V amons mm H mH ss-ono N NH H +N m H.o- e.m onm.H m.mH- o.mH ons.o om.H om H.cm H.e~ +~ e.o- m.m OHx~.m m.-- o.eH onmo.H oH.H um H.es H.em +~ a m.mH con.H ow- H.mH OHx~.~ so am- ~.o- m.oH Cme.m ~.m- e.MH onm.o oo.H no H.sm H.eo +~ fieonm fl$me mm H +s o I. o o I. o o c m m m o o . oons a HH m NH H.em nH No H + 2 H a 5.0 . Hoe. .233 . com fl 5.0 :33. Hoos— . oom < sown—mo o H- o H- o H- o H- o H-n m +m< +m< u— +m< +=< o— H w m o o o o o Aa.N.NV oaom mo coaumaoommaa ocm coauoeuom How muouosmumm coaum>auo< ocm mucmumcoo oumm .Uomm um cosmonauu .m oHome 78 Notes to Table 7. aCrystallographic ionic radius, from Table 6. bRate constant for cryptate dissociation, acid-independent pathway. cActivation enthalpy for dissociation. dActivation entropy for dissociation. eRate constant for cryptate formation. f Activation enthalpy for cryptate formation. gActivation entropy for cryptate formation. Data Sources: hB. G. Cox and H. Schneider, to be published; determined at u~10"3 - 10‘2. iRef. 27. jRef. 24. kRef. 26. 1From Table 3. mCalculated from the kinetic parameters for cryptate dissociation combined with the thermodynamic data in Table 6. 79 AmV Am.NHV Em.~ cm- w.mH HMIOHxH.H mm.o +mnm ' o o In 0 o o D AoH V aH mV emOme e 0H m mH am-ono m NH H +N m mHu m.o OHxn ml m.o~ IOHx~.~ so an m.mHI ~.o Ome m.m| o.HN Owa.H om.H mm H.es H.cm +~ Con OHxH on oel mHI «.5 Oon.H NHI n.hH Oon.H as se- a HHI m m nHuo< one .osmw on successes munmumnoo comm .m «Home 80 Netes to Table 8. aCrystallographic ionic radius, from Table 6. bRate constant for cryptate dissociation, acid-independent pathway. cActivation enthalpy for dissociation. dActivation entrOpy for dissociation. eRate constant for cryptate formation. f Activation enthalpy for cryptate formation. 8Activation entropy for cryptate formation. Date Sources hB. c. Cox and H. Schneider, to be published; determined at u ~10"3 - 10'”2 . 1Ref. 27. jRef. 24. "Ref. 26. 1 From Table 3. mCalculated by combining the kinetic parameters for cryptate dissociation with the thermodynamic data in Table 6. “Ref. 22. oRef. 3. 81 is likely that these behavioral differences between Eu3+ and the lower charged cations result chiefly from the influences of ionic rather than covalent bonding. The first possibility to be considered in interpreting such variations in the complexation kinetics is that they may be connected with the rates of substitution of water molecules in the coordina- tion sphere of the cations. However, the "characteristic rates" of water substitution (105) for tripositive lanthanides, heavier alkaline earth, and alkali metal cations are all extremely fast (107-108, 9, and ca. 109 sec-1, respectively (105). It has been sug- 108-10 gested (25) that the rate contents k.c for complexation of cryptands with alkali metal and alkaline earth cations that are markedly below these water substitution rates may be due partly to the need for the complexing cation to bind with the endb-endb form of the cryptand, whereas the exo-exo conformation could be the predominant form in solution. However, since the 106- to 107-fold difference in kc between Na+ and Eu3+ , for example, is much greater than the corres- ponding variation in water substitution rates, it is likely that an additional factor is chiefly responsible at least for the large energy barriers observed in the complexation of multicharge cations. It is plausible that an important component of these energy barriers arises from the inability of the relatively rigid ligand to achieve a smooth stepwise replacement of water molecules in the metal coordination sphere by the cryptand coordinating groups, so that the loss of sol- vating water molecules may not be immediately compensated by the formation of cation-cryptand bonds (3,26). 82 As noted previously (3,25,26), the large changes in cryptate thermodynamic stability induced by variations in the radii of uni- valent or divalent cations (Table 6) are chiefly reflected in k.d rather than kc (Tables 7 and 8), suggesting that the transition states more closely resemble the separated cation and cryptand rather than the cryptate (3,25,26,28). This latter result can be reconciled with the larger sensitivity of kc to cation charge by noting that the hydration energies of simple cations are much more sensitive to ionic charge than to size (103). Therefore any partial desolvation of the cation required to reach the transition state from the separated reactants is expected to result in a greater dependence of the activation barrier on the charge rather than the size of the cation. Further insight into the factors influencing these reactivity patterns can be obtained by inspecting the corresponding activation parameters (Tables 7 and 8). By and large, the marked variations in As; seen for the univalent and divalent cryptates as the cation size and charge varies (Table 6) are primarily reflected in the activation 1? d. activation entropies for complexation As: are.small and negative entrOpies for the dissociation reactions AS The majority of the (ca. -5 to -10 e.u.). Since an activation entropy of ca. -10 to -15 e.u. is predicted on statistical grounds from the loss of trans- lational entropy ASt attending such association reactions (106), the values of ASc support the above contention that no drastic changes in the solvent structure surrounding the cation are required in order to form the transition state from the separated reactants. 83 In contrast, however, the large differences in As; between Eu(2.2.1)3+ and Eu(2.2.l)2+, and between Eu(2.2.2)3+ and Eu(2.2.2)2+ are chiefly reflected in As: rather than as: (Tables 7 and 8). Consequently, the markedly smaller values of kc for the Eu(III) cryptates compared to Eu(II) and other divalent cryptates arise from substantially larger enthalpic barriers AH: (Tables 7 and 8). The differences in activation entropy suggest that the solvent surrounding the Eu3+ cation in the transition state is substantially less "ordered" than around the uncomplexed cation. Indeed, the ionic entropy S9 of E113+ is large and negative (107), reflecting its large solvent "structure-making" ability (108). A good deal of this exten- sive hydration shell may be required to be removed in order to form the transition state, resulting in the observed large positive values of AH: as well as ASI. c It has been pointed out that the significant acid catalyses observed for the dissociation of a number of metal cryptates suggest that it is often necessary to alter the cryptate conformation from the stable endb-endb form to one where at least one of the two bridgehead nitrogens has its lone pair pointing outwards (exo-endb or exo-exo conformation) prior to or within the transition state (27). The significant acid catalyses of the dissociation of the Eu(III) and Eu(II) cryptates (Table 4) indicate that such a mechanism is also applicable to these systems. Inasmuch as any clear trends can be discerned in the ratios kH/kd of the rate constants for the acid- catalyzed and uncatalyzed pathways, it appears from the data in Table 4 and reference 27 that kH/kd increases as the charge/radius 84 ratio of the cation increases and as the thermodynamic (or kinetic) stability of the cryptate decreases. The ratio kH/kd also is gener- ally larger for (2.2.2) compared with (2.2.1) cryptates. The role of the proton in accelerating the dissociation reaction may be to aid the endb-endb to exo-endb conformational change by forming an incipient nitrogen-proton bond as the nitrogen-metal bond is broken, or it may simply provide a more favorable dissociation pathway once this conformational change has occurred (27). The smaller or undetectable acid catalysis in the more stable cryptates may arise from the greater difficulty of accomplishing this conformation change due to their tendency to be less flexible than cryptates having cavities that are markedly larger than the encapsulated cations. The very marked acceleration of the dissociation rates of tri- valent lanthanide cryptates by OH- or F- is not surprising in view of the strong binding of these anions to the cryptates. These anions could lower the energy barrier to dissociation by distorting the cryptate conformation and enlarging the space between the polyether strands so that the encapsulated cation can exit the cavity more easily. Additionally, the bound anions will act to lower the effec- tive charge of the assembly, thus reducing the extent of the solva- tion changes which must occur in order for the cation to be removed from the cryptate cavity. It is interesting to note that these sub- stantial accelerations of the cryptate dissociation rates occur in spite of the considerable thermodynamic stabilization of the cryptates by both fluoride and hydroxide (Table 2). CHAPTER IV REACTION ENTROPIES 0F TRANSITION METAL REDOX COUPLES 8S A. Introduction While the ionic entropies of a number of transition metal com- plexes have been determined (see Historical Background), no systematic study has been conducted on the dependence of the entrOpies on the nature of the metal ion or on the ligands coordinated to the cation. Ru(III)/(II) couples can be used to explore the ligand effects since Ru(III) and Ru(II) complexes containing a variety of ligands are known. As stated previously, the substitutional inertness of Ru(III) and Ru(II) allows electrochemical determinations of the reaction entropies of these systems. The influence of the central metal ion on the ionic entropy of its complexes can be investigated by obtain- ing the reaction entrOpies for a series of redox couples having the same ligand structure and charge type. A.large number of M(III)/M(II) couples is available so that complexes of this charge type will be used to investigate the metal dependence. A wide variation in metal ions is not possible with.most unidentate ligands due to the instability of many of these complexes with regard to aquation. Fortunately, many complexes containing chelating ligands should be sufficiently stable to maintain their structural integrity in both the oxidized and the reduced forms. Therefore the employment of chelating systems will allow one to achieve the desired diversity in the nature of the complexed metal ion. 86 87 B. Determination of Reaction Entropies The reaction entropy has already been defined in Chapter I but it is useful to present this definition more formally. One can con- sider the reaction M¢IIL'L" + e-(metal electrode) ‘+ MIIL'L" (1) m'n m n in which.M is a metal which has trivalent and divalent oxidation states, L' and L" are neutral or anionic ligands, and the number of L' and L" coordinated to M is given by m and n, respectively. The reaction entropy A8;c of the MIII L'L"/MIIL'L" redox couple can then m n m n be defined as the standard entrOpy change for reaction (1). This definition can also be stated as O . -0 - -0 ASrc 311 S111 (2) where §° and §° are the absolute ionic entropies of MIIL'L" and II III m n MlllhéLg, respectively. The As;c definition can be expanded to include other classes of redox couples but this work will only deal with those of the M(III)/M(II) type. Reaction (1) is only a half of a complete electrochemical cell reaction. Consequently, any determination of the AS°for this reac- tion must involve some sort of extrathermodynamic assumption. Several of these have been proposed, and some reliable methods for the quantitative estimation of ionic entropies and As;c values have been developed from these assumptions. Criss and Soloman have assembled a helpful summary of the various procedures (40). 88 Of the several methods-of determining ASrc which are available, the most convenient from the viewpoint of the present study is the one which utilizes a nonisothermal cell (109,110). This is basically an electrochemical cell in which the reference and work- ing compartments are.maintained at different temperatures. One nonisothermal arrangement which was frequently employed in the present work can be written as: ' CuIHgIHgZClz, A KCl(sat) I I KC1(3.5M) IKC1(3.5M_) I luau) ,M(II) IHg IHgI Cu B c n The temperature in regions AB and CD was maintained at the ambient value, while that of region BC (i.e., the working compartment) was ni varied, usually between 3° and 60°C. The formal potential Ef across this nonisothermal cell was determined as a fuzcgion of temperature. The temperature coefficient of this potential :T can be divided into three components cuff11 Pill d¢tc do:l 71'— ' dT + 71'1— + “d? (3) where ¢tlj is the Galvani potential difference across the thermal liquid junction within the KCl salt bridge (point B in the cell diagram), ¢tc is the "thermocouple" potential difference between the hot and cold sections of the mercury working electrode, and 0: is the Galvani metal-solution potential at the working electrode. The important quantity in the present content is do:/dT since 0 m ASrc a F(d¢f/dT) (4) 89 Therefore one can obtain A8;c values from.measurements of the formal potential if the temperature coefficients of 0 tlc and ¢tc are known or can be estimated. Absolute values of the Thomsen coefficient, dotc/dT, are known for several metals and are usually found to be on the order of a few microvolts per degree. Mercury and platinum, which served as the working electrode materials in this study, have Thomsen coefficients of about 14 and 6 uV/deg, respectively, over the 0°-100°C tenperature range (111). The experimental dEgildT values are usually 1-2 mV/deg so it is clear that the contribution of the Thomsen coefficient to the temperature coefficient of En1 is essentially negligible. f Relative rather than absolute values of d0 ldT are thermo- tlj dynamically accessible but there is good evidence to indicate that this quantity is not greater than 50 uV/deg for most aqueous electro- lytes (109,110). Much larger values are found with strongly acidic or basic media (llOb), in analogy to the considerable isothermal liquid junction potentials which can arise in such media. The use of concentrated aqueous KCl in the region of the thermal gradient (in effect, the employment of a "nonisothermal salt’bridge") has been suggested by de Bethune et a1. (llOb) as a way to minimize d0 IdT. tlc Although reservations have been expressed concerning the exact sound- ness of this assumption (llOc), there is little question that dotlj/dT is at most 20 uV/deg and is quite likely to be considerably smaller. htremely accurate measurenents of ASI‘:c will definitely be hampered by the uncertainty in do /dT, but this uncertainty will be tlj essentially negligible within the context of the present study, in 90 which the precision of the dErflildT determinations was only :50 uV/deg. An uncertainty in As;c of i0.5-l e.u. arises from this thermal liquid junction problem but this is quite small compared to the measured Asgc values, which can range up to 50 e.u. Furthermore, comparisons of the As;c values of various systems will be unaffected by the dotlj/dT effect since the value of this last quantity will depend on the composition of the salt bridge and not on the individual redox couple under study. The AS° values for homogeneous redox reactions, which can be obtained from the difference in As;c values of the redox couples involved, will be free of the above effects since the uncer- tainties in As;c will cancel when this difference is taken. On the basis of the preceding discussion it can be concluded that, for the purposes of the present study: 0 ni Asrc - F(d1:f MT) (5) The formal potentials which are needed for determinations of As;c can be conveniently obtained from measurements of Elh’ the (polaro- graphic) half-wave potential. For a reversible cyclic voltammogram, 35513 the average of the cathodic and anodic peak potentials (92b). A simple relationship exists between E and E1/2 (93,112): f 1 21,2 - s + (RT/nF) 1n (DH/D ’2 (6) f III) where DH and DIII are the diffusion coefficients of the reduced and oxidized species, respectively. The DII/D ratio is usually close III to unity so that E55 differs from E by only 2-3 mV. In the context f of the present work, the temperature coefficient of the 91 (RT/nF)ln(D >72 term must be negligible so that dEfitldT can be II/DIII / equated to dEgildT. The validity of this assumption was confirmed by using the limiting polarographic and cyclic voltammetric peak currents of the oxidized and reduced forms of the representative - 3+/ 2+ 3+/ 2+ E“ 3)6~ as a function of temperature (113). These determinations allow one and Ru(NH , to obtain values of D and D couples, II III to estimate that an error of less than 0.5-1 e.u. in As;c stems from the equation of the temperature dependences of 3:1 and Efii. Therefore, one is justified in writing 0 ni Asrc F(dE5&/dT) (7) The procedure described here yields As;c values which are absolute entropy differences for redox couples. Other reaction entropy measure- measurements and scales are found in the literature. Unfortunately, little has been done to reconcile the conflicting conventions which exist. One pOpular scale stems from the determination of reaction entropies from the temperature dependence of the cell potential of an isothermal cell (114); i.e., one in which no temperature gradient exists between the working and reference electrode and in which the temperature of the entire cell is varied. Reaction entropies for single redox couples are then obtained by assuming As;c - 0 for the reaction H+‘+ e ’5 )5 H In effect, a relative scale is established 2. for As;c. Reaction entropies on the absolute scale can be obtained by adding 21 e.u. to the relative values (40,114). (This conversion is derived by recalling the estimated entropy of 5 e.u. for H+ in aqueous solution (40).and by combining this with an entropy value of 92 31.2 e.u. for hydrogen gas at 25°C (115) to compute a value of As;c - 21 e.u. (on the absolute scale) for the H+IH2 couple.) As noted in Chapter I, a relative scale of ionic entrOpies can be set up by arbitrarily assigning a value of zero to the ionic entrOpy of hydrogen ion. Such assumptions will not affect the value of the reaction entrOpy since they will cancel when the difference between ionic entropies is calculated. However, care must be taken to avoid mixing values from the relative and absolute scales when taking such a difference. Both scales are used in the literature, and in some papers it is not obvious which scale is being employed. Furthermore, As;c values from.older papers often refer to oxidation processes. Such values will have the apposite sign to those in this work where the reaction entropy is defined in equations (1) and (2) in terms of reduction reactions. C. Practical Aspects of Nonisothermal Cells As shown in the preceding section, the nonisothermal salt bridge is a crucial component of the nonisothermal cell. In the present experiments, 3.5 M_KC1 was employed in the salt bridge for most experiments. The solution in the bridge must contain cations and anions of similar, if not identical, mobilities (116). This require- ment is met by K+ and Cl_ (116,117). The solubility of KCl is slightly greater than 3.5 M even at 2°C, the lowest temperature used in this study, so no problems with precipitation of KCl are encountered. The composition of the salt bridge can be important under some 93 conditions. With experiments performed in neutral supporting electro- lytes, a negligible change in dEgi/dT was found when the supporting electrolyte was used in the salt bridge in place of the 3.5 M KCl. An appreciable alteration in the value of dB: the analogous comparison was made with acidic supporting electrolytes i/dT was observed when (i.e. where [H+] >‘52 of the total ionic strength). In these par- ticular cases, 3.5 M KCl was enployed to minimize the isothermal junction potential between the salt bridge and the acidic supporting electrolyte. For experiments conducted in concentrated perchlorate electroq. lytes, one cannot use 3.5 M KCl in the salt bridge since the precipi- tation of the sparingly soluble salt, KClO in the liquid junction 49 will give rise to spurious potentials. This problem can be avoided by employing 3 M NH4C1 in the salt bridge. The mobilities of NH: and C1. are quite similar (117), so NH Cl is a good replacement for KCl. 4 Some attention.must also be devoted to the actual physical dimen- sions of the salt bridge. The electrochemical cell was constructed so that the temperature drop within the nonisothermal salt bridge was confined to a short distance (<1 cm) within a glass tube having an internal diameter of about 0.8 cm. These features prevented the development of any significant concentration gradients due to thermal diffusion (the Soret effect). Considerably larger values of dotlj/dT would result from the presence of a Soret effect (109,110b) but the lack of any significant thermal-diffusion effect was demonstrated by the fact that cell potentials were found to be stable to within 1 mV'even after several hours under nonisothermal conditions. D. Results An orderly presentation of the results is facilitated by dividing them into three categories on the basis of the ligand structure and type of central metal ion. This arrangement will also allow one to make comparisons within and between classifications to determine the factors which influence the reaction entrOpy. 1 . Acme Couples The reaction entropy As;c of six aquo couples of the type 3+/ 2+ , 3+/2+ 3+] 2+ 3+/ 2+ M(OH2)n were studied. Cr(OH2)6 , V(OH2)6 , Fe(OH2)6 , Ru(OH2)2+/2+, Eu(OHz):+/2+, and Yb(OH2):+/2+. These systems were chosen because their formal potentials allow them to be conveniently studied at mercury or platinum electrodes, and they exhibit sub- stantial variations in the electronic structure of the central metal ions. Previous determinations of As;c for these couples are sparse. The absolute ionic entropies of Fe3+ and Fe2+ have been determined (41), and an estimate of Asgc for Ril(032)g+/2+ has been reported from electrochemical measurements (118). 3+/ 2+ 3+/ 2+ Most data for the Cr and Eu couples in perchlorate media were obtained using potentiometry because the small heterogeneous electron-transfer rates for these systems resulted in distinctly irreversible cyclic voltammograms. However, in sodium p-toluene- sulfonate (NapTS) media, the strong specific adsorption of p-toluenesulfonate anions resulted in almost reversible cyclic voltammograms for the Eu3+/2+ couple. Accurate values of E1, could 2 still be obtained from such "quasi-reversible" voltammograms in the usual way provided that the cathodic-anodic peak separation lies in 94 95 the range 57 to ca. 90 mV (93). For the remaining aquo couples, essentially reversible or quasi-reversible cyclic voltammograms were obtained, at least after the addition of small quantities of NapTS. The resulting values of As;c are listed in Table 9, together with other pertinent information. It is seen that some limited dependence of As;c upon the nature of the metal ion is observed. The dependence 3+/ 2+ of As;c upon ionic strength was investigated for Eu and was found to be small (Table 9). Good agreement between the earlier and present determinations are found for Fe(OHz):é+/2+ (Table 9), but a large qualitative discrepancy is seen for Ru(OH2)g+/2+. A possible reason for this disparity is that the isothermal cell measurements of ref.118. were complicated by an unknown temperature dependence of the electrode potential of the glass reference electrode used in that study. 2. Rutheniun(111)/(II) and Osmium(III)/ (II) Couples Containing Monodentate Ligands The reaction entropies of twelve Ru(III)/(II) couples containing ammine, aquo, and simple anionic ligands were evaluated. These systems were selected in order to scrutinize the effects of replac- ing ammine by aquo ligands and of changing the charge type of the couple resulting from substitution of the ammine and aquo ligands by simple anions. The results are summarized in Table 10. For all.of these systems, the heterogeneous electron-transfer rates were suf- ficiently rapid so that the cyclic voltammograms were essentially reversible even at the highest sweep rates (100 V s-l). However, 11 2+/+ the relative lability of the Ru state for the Ru(NH3)5Cl and 96 I- I. I a II o o N HHmm V bosom 8 m 3 H2. m an o +~ \ +1 :85 a o .- c a. II II \I\ Q 0 Home 8 m 2.: us a V use an o a +N\+nH .18.; Homes oo-m oao- o3 mnVaeHusz mH one oo-m oao- u? rnVHmaeoz mic . I I n on .I . n N on as 8 m was on” m V H8. 2 :8 o H +~H+mH more N” cl II ..o I. Q q II o o N o 8 m as 8 H H e V 303 an o H +~H+nH no; A ”Q m I” o I. Q Q I o 0 N 0 r V no 8 8... SHHHHVoHoHH one n+2+mH =8 e o I o N I I n o u use 8 n one 8 m V on. z E H +~H+mH :8 o and.” .00 >6 a cums omens case 3.383on oHesoo .oamH .moHenoo wooom Ha +N\+m Ammovz mnOHum> pom moHnouunm noHuooom .m oHoma 97 Notes to Table 9. aReversible half-wave potential in mV, determined at 25°C against a KSCE held at ambient temperature (23 i 0.3°C); related to E by f equation (7). bReaction entropy of redox couple in e.u. (cal K.1 mole-1) at 25°C; determined by using equation (6). Experimental precision estimated to be 11 e.u., accuracy within 1-2 e.u. (see text). values in parentheses are from literature sources. cDetermined using a combination of potentiometry and dc polarography. l dDetermined using cyclic voltammetry (sweep rates: 50 - 500 mV s- ). e355 found to be unaffected by pH variations of at least one unit around the pH value given. fDetermined using HMDE. gDetermined using Pt electrode. hCalculated by taking the difference in the relative ionic entrOpies given in ref. 41. 1Calculated from isothermal cell data given in ref. 118 by noting that Asgc .- 21 e.u. for the reaction H+ + e- e 1le2 (40,114). i pTS - p-toulene sulfonate. kDetermined by P. D. Tyma. 98 s.os H oH ra-H Hanan.ussmVonm soHuoz.mH oH+HHoon=zVse-ouo H.n~ H oH ms-H AnorHVmoH HOHonz_mw.o +H+HHQHHH=zVsm on.o H mH oo-m HmHHHVoaH cans ms.o om.o H «H oo-m omH oaas_muo.o +x+~wozmHH=zVse uH H NH oo-m me- ozooommo mH.o H +H\+msanamszsm on H o om-m Hmooonmo mooz_mw.o +H+Hmomam=anm oH H om oo-m HeooH.mosHmeH mene_mH.o +~H+Ma~moVsAmsz:m-uoo HHm.aHVoH H mm oo-m HessH.mmaHV~oH mooomm0.m~.o +HH+MeomHmszsm om.o H mH mo-m HHoHoHVoss ozooomno.mwo.o +~H+mhmszno on.o H sH oo-H er ozooomao_mn.o om.o H n.oH oo-m onH ozooomno_mu.o HHaVo.om.o H oH oo-mH naH soHusz.mH.o on.o H HH oo-HH HeerH.nmaHerH some me.o om.o H sH oo-m HHH seas moo.o +~H+MHHansm A. 9.3 omm< Us . owned 95 “Hal 3.30.50on oHnnoo A . neon. n mcHdHousou ooHnsoo noose HHHHHHHVoo one HHHHHHHVne .oncanH ououcoana mnOHuo> you moHoouunm noHuonmm .CH 038. 99 - I . - . N a N I area + 2 mm s 8H mean an e or. Ho A more one - n I m N on H mm 3 m 3H 2. m 2H +\+~Hu A more ram-Vatom H on om-m romVoH menu mic AEMHHHSVSH on a.c.mv omo 0o .ownwm 35 WT oumHowuumHu oHonoo A .93“. n A3363 .3 Home 100 Notes to Table 10. aReversible half-wave potential in mv vs. ambient KSCE; determined at 25°C by cyclic voltammetry at a HMDE (except for Ru(NH3)5py3+/2+ in which a Pt electrode was used). See notes to Table 9 for further details. values in parentheses are from the indicated literature sources and correspond to comparable experimental conditions. bReaction entrOpy of redox couple in e.u. at 25°C (see notes to Table 9). Stated precision was estimated from the scatter of experimental points in the vicinity of 25°C. Values in parentheses are from the indicated literature sources. cDetermined using cyclic voltammetry sweep rates in the range 50-500 mV s-l. dDetermined using sweep rates in the range 1-100 V 3‘1. epTS - p-toluenesulfonate. fpy - pyridine. 8Reference 52. hReference 54. 1Reference 56. 1Reference 119. kReference 118. lDetermined by K. L. Guyer. mDetermined by P. D. Tyma. 101 Ru(NH3)l‘Cl-;/o couples necessitated the use of large sweep rates II (1-100 V sec-1) in order to avoid significant aquation of Ru during the potential scan. Estimates of As;c have previously been obtained i+l2+ §+IZ+ and Ru(en)§+/2+ from the tempera- ture dependence of the equilibrium constants for reduction of the for Ru(NH3) , Ru(NH3)50H Ru(III) complexes by Npiz coupled with an estimate of As;c for the 4+/ 3+ Npaq couple (119). While reasonable agreement between the present and earlier determinations is found for Ru(en)§+/2+, substantial dif- ferences are seen for the other two systems (Table 10). These dis- crepancies may arise from systematic errors in the kinetic analysis employed to determine the equilibrium constants in ref. 119 when these quantities are much larger than unity. Significant differences are also seen between present and earlier determinations of E for 72 some Eu(III)/(II) couples (Table 10). In view of the care taken to minimize liquid junction potentials in the present work, these dif- ferences probably arise chiefly from the presence of such potentials in the earlier work combined with the slightly different thermal con- ditions employed here. Two main trends are seen upon inspecting the data given in Table 10. Firstly, the stepwise replacement of ammonia by aquo ligands results in large increases in As;c. Secondly, the substi- tution by anionic ligands results in significant decreases in us;c. A value of A8;c for Os(NH3)2+/2+ is also given in Table 10 for com- parison with Ru(NH3)g+/2+. The reaction entropies for these two systems containing the same ligands are in close agreement. 102 3. Redox Couples Containing Chelating Ligands In contrast to redox couples containing only simple unidentate ligands as considered above, a significant quantity of information has been gathered previously on the reaction entropies of couples containing chelating ligands. One reason is that such complexes are often sufficiently stable so that even labile oxidation states can remain in the complexed form in the presence of small stoichiometric excesses of the chelating ligands. Consequently it becomes feasible to determine values of As;c for redox pairs such as Co(III)/(II) when bound to some chelating ligands. The lability coupled with the weak complexing ability of the Co(II) state precludes such studies of Co(III)/(II) couples containing only unidentate ligands. The simplest example is Co(en)3+/2+ 3 3+/ 2+ with Ru(en)3 . The relevant data for these couples are given in , which provides an interesting comparison Table 11. It is seen that the values of A8;c for these two redox couples are strikingly different. The large value of As;c for 3+/ 2+ 3 smaller values seen for other amine complexes (Table 11), but is Co(en) (37 e.u.) is surprising in view of the markedly in accord with an earlier determination (120). Consequently, we decided to evaluate As;c for some other related Co(III)/(II) couples. The tris-l,lO-phenanthroline and tris-2,2'-pipyridine Fe(III)/(II) and Ru(III)/(II) couples have previously been found to exhibit values of As;c that are close to zero (51,121). The present determinations of As;c for Fe(phen)§+/2+ 3+/ 2+ and Fe(bpy)3 are in close agreement with these earlier values. 103 m 88 no man + H H m.¢H oolm mwu mzooommo meo.o +N\+mnoroonmsznm w.o H H m.~H oolm «mm ozooommo.mH.o +~\+mhooeAmsznm wao H H o oo-m Hos once HHH +HH+MHxnoVHHm=zVse-nso w.o rooomeo maoH H H no oo-m «so .ozooomeo mic HHxaoVHHmszsH-uso To +~\+m - m I. m an. I "1. . sham HH fo 8 H + H on m fa 2:333 58 no 5 c we +HH+HH 5 a sec mama + - .I m I:- I o ".1. o sham HHJHo V H + N 3 .H f8 o 389% Hon :3 o in +Q+HH oVue core mafl + I. m "1. I I o "1. o a m HHJHo 8 N + n no .H Heme o moonS Us :3 o no +N\+m?uc V n so maom + I a .V I. m ens. HV H + R oo-m H as nos-V8.7 332 E 338 ho ho 0.0 +~\+m . - I - o l . n no HHHHV m o + H 8 m 8 no. 2H o oJ +H\+HH Van A.n.oV ouom< .oo .ownnx A>nV mam ouhHoHuooHu oHonoo .monmeH wnHumHonu wanHmunoo mmHonoo aHHv\AHHHVz mnoHuo> How moHoouunm noHuomom .HH oHomh 104 N H m- Has-H NH .683 mHo NHmnmoVue Ho 3+ I Q I - - - . e N + 2 on N Acorn V can 39.2 H: o :6 +N\+NH 938 sec mamN + l I I o M %Q HaHNV m + NN 3 m 2 H8. :3 o no +N\+mH 38 core mamN + I I I. . m n o HamHV m + NN 3 m 3H H8. :8 c on. +N\+mH on V8 H H no oo-m N3 mane mH + NHxnbVNHNeoVé-orneu , To N\+m H H no co-m car was: HHH + NercVNHNmoVas-nsu To N\+m . . on . N\ A a CV: owe .oo ownmm A>EV m Hm mumHouuoon oHonou .oeoa A.unooV .HH «Hana 105 Notes to Table 11. aReversible half-wave potential in mV vs. ambient KSCE; determined at 25°C by cyclic voltammetry using sweep rates of 50-500 mV 5-1 (see notes to Table 9 for more details). values in parentheses are from literature sources and determined at the indicated ionic strength u. bReaction entrOpy of redox couple at 25°C in e.u. (see notes to Table 9 for further details). Quantities in parentheses are literature values determined at the indicated ionic strength. No 11 value is given when the literature study and the present study employed the same ionic strength. cDetermined using HMDE. dDetermined using platinum and glassy carbon electrodes. een = ethylenediamine. fphen - 1,10-phenanthroline. 8bpy 8 2,2'-bipyridine. hsep - l,3,6,8,10,l3,16,l9-octaazabicyclo(6.6.6)eicosane. The trivial name for this ligand is sepulchrate (see reference 88). 1Reference 119. 1Reference 120. kReference 121. 1Reference 51. mCalculated from isothermal cell data (A. Claus and V. Crescenzi, quoted Table V of M. Chou, C. Creutz, and N. Sutin, J. Am. Chem. Soc., 22, 5615 (1977). nReference 88. 106 Relatively small values of As;c are also maintained for cis- and trans-Ru(OH2)2(bpy)2+/2+ (Table ll). However, the measured values of As;c for Co(phen)§+/2+ and Co(bpy)?"2+ are both substan- tially different from zero (ca. +25 e.u., Table 11). It therefore again appears that such Co(III)/(II) amine couples exhibit "anomalously" large As;c values. On the other hand, As;c for the "capped ethylenediamine" Co(sep)3+/2+ couple (88) is much smaller than for Co(en)§+/2+ (Table 11). E. Discussion l. ILigand Effects on Reaction Entropies It has been known for some time that the entropies of simple aquo cations can be correlated with surprising success using empirical relations involving the ionic charges and radii (41,45). Using the Latimer-Powell relation (41) and known crystallographic radii (l8), As;c for the first-row transition metal couples in Table 9 are cal- culated to be ca. 45 e.u., in good agreement with the experimental results. However, the smaller values observed for Ru(OHz)‘2+/2+ and )3+/2+ 3+] 2+ 2 6 n and Yb(OH V(0H (36 and 37 e.u.) and the larger values for Eu(OHZ) ) 3+] 2+ 2 n classical Born equation can only yield predicted values of As;c in (48 e.u.) are not predicted by this relation. The reasonable agreement with these results by inserting the radii of the bare cations (45), rather than the radii including the coordinated water molecules which intuitively seem more appropriate, especially for transition metal cations. These large experimental values of As;c are probably a consequence of the release of water molecules 107 surrounding the primary coordination sphere that are strongly orientated ("frozen") in the tripositive oxidation state (47,48). Indeed, such solvent structuring effects provide the most widely accepted and intuitively reasonable explanation for the wide varia- tions found in As;c (47-49,51). In this connection, it is interesting to note that simple M(III)/(II) ammine couples exhibit values of A3;c which are substan- tially less than for the corresponding aquo couples (Tables 9 and 10). 3+] 2+ 3+] 2+ 6 6 ’ and other hexaaquo couples, Astoc 2 36 e.u. Since the Thus, for both Ru(NH3) ) 3+] 2+ 2 6 size, shape, and electrostatic prOperties of ammonia and water ligands and Os(NH3) AS° ~18 e.u., whereas rc for Ru(OH are comparable (122), these differences suggest that rather specific interactions between the coordinated and surrounding water molecules are responsible for the large values of As;c observed for the aquo systems. It seems reasonable that the large degree of solvent order- ing around tripositive, compared to dipositive, aquo cations arises partly from.the ability of the relatively acidic aquo protons to form hydrogen bonds with surrounding water molecules (47,48). The weakly acidic ammine protons presumably have a much lower tendency to aid the central cationic charge in orienting solvating water molecules in this manner. These results are not unexpected on the basis of the empirical entropy correlation of George et al., which also indicates that the entrOpies of ammine complexes decrease less with increasing cationic charge compared with aquo systems (43,44). The much lower "structure-making" ability of tripositive ammine, compared with aquo, complexes is also borne out by the much smaller effective hydrated 108 radii for ion transport that are observed for the former complexes (123). The differing magnitudes in the electrostatic double-layer effects observed for the electroreduction of Cr(III) aquo and ammine complexes suggest that such differences in the extent of hydration also survive within the electrode-solution interfacial region (124). 47.1 re where re is the effective radius of the ion (45), and the radius of )3+/2+ 3 6 couples AS]:c “l6 e.u. The close agreement between the Born prediction 3+] 2+ 3+]2+ 6 6 (Table 10) indicates that there is no extensive solvent ordering Since the Born relation predicts for +3l+2 couples that As;c - e.u. M(NH is ca. 3 A (122), then this relation predicts for the and the experimental results for Ru(NHB) and Os(NH3) around these species even in the tripositive state, although dielec- tric saturation effects (46) may complicate the application of this simple model even to such substitutionally inert complexes. The substitution of ammine by aquo ligands in the series of couples 3+]2+ O Ru(NH3)6_x(OH2)x is accompanied by sizable increases in Asrc’ especially for x - 1 (Table 10), indicating that substantial solvent structuring can occur even around an isolated aquo ligand. The experimental values of As;c for both Eu(OH2):+/2+ (Table 9) and Ru(NH3)g+/2'+ (Table 10) depend only slightly upon the total ionic strength, confirming that this quantity is chiefly a consequence of ion-solvent, rather than ion-ion interactions. The small decreases in As;c at ionic strengths approaching unity are probably due in part to ion-pairing in the tripositive oxidation state, so that the effec- tive ionic charge of this state is less than +3. 109 The substitution of ammine or aquo ligands by simple anions in Ru(III)/(II) couples consistently results in substantial decreases in AS££ (Table 10). This effect could arise for at least two reasons. Firstly, the reduction in the net ionic charges is always expected to lower 63;; on the basis of the classical Born model since As;c is predicted to depend on the difference of the squares of the ionic charges on the two ions forming the redox couple. Secondly, the replacement of hydrogen-bonding ligands by electronegative ligands should decrease the extent of solvent structuring around these cations. To varying degrees, these two factors are probably respon- sible for the observed behavior of the chloro and isothiocyanato complexes in Table 10. The more dramatic decrease in As;c for 2+]+ 3+] 2+ 2 be attributed to the different hydrogen bonding characteristics of Ru(NH3)SOH compared with Ru(NH3)SOH (Table 10) can reasonably OH- and 0H2. Thus OH- can hydrogen bond to surrounding water mole- cules via the oxygen atom, which should be favored on electrostatic grounds by coordination to Ru(II) rather than Ru(III). In contrast, hydrogen bonding involving aquo ligands will be electrostatically favored by coordination to Ru(III) rather than Ru(II) since the hydrogen atoms on the ligands will participate in the bonding. Although the data for the aquo couples in Table 9 indicate that some dependence of As;c upon the electronic structure of the central metal cation is to be expected, part of the variations could arise from differences in the number of aquo ligands that are bound to the central metal ion. Thus n is undoubtedly greater than six for 110 3+]2+ 3+] 2+ 2)n and Yb(0H2)n hydrogen bonds with the surrounding solvent and may account for the Eu(OH which can result in a greater number of expecially large values of As;c that are observed for these couples. The reduction of V(III) to V(II) could be accompanied by an increase in the number of bound water molecules (including those held by hydrogen bonding). Such a change is sterically reasonable in view of the large ionic radius of v2+ (r “‘0.8 A (11)) and would result in a smaller value of As;c, as is observed. Similarly, Cr(II) is highly Jahn-Teller distorted (125), and the axial ligands in Cr(OH2)§+ are very weakly held. As a consequence, the conversion of Cr(III) to Cr(II) is accompanied by a decrease in the number of bound water molecules. This could be responsible in part for the relatively large value of As;c. Differences in the polarizing power of the various cations may also be a factor. The As;c values for Ru(III)/(II) couples containing simple unidentate ligands given in Table 10 could provide workable estimates of As;c for other couples with the same ligand constitution and charge type. However, it is apparent from the data given in Table 11 that this assumption may not be reasonable for complexes that contain chelating ligands. Thus the values of As;c for Co(III)/(II) complexes containing ethylenediamine, o-phenanthro- line, and bipyridine ligands are about 25 e.u. larger than for corresponding complexes involving other metal cations. These results are both surprising and somewhat puzzling. The behavioral difference 3+] 2+ 3+] 2+ 3 3 iriligand conformation that are known for the trivalent complexes and Co(en) might be explained by differences 'between Ru(en) (2126) giving rise to differences in the surrounding water structure. 111 However, this explanation seems less plausible for the much larger and structurally more open phenanthroline and bipyridine complexes. The values of As;c close to zero that were previously found for other phenanthroline and bipyridine couples have been explained by the supposed ability of these aromatic ligands to "shield" the central metal ion from the surrounding solvent (51,121). The contrasting behavior of the corresponding Co(III)/(II) couples clouds this explanation somewhat. Certainly a common feature of these Co(III)/(II) couples is that the change in electronic structure (t28)6-+ (t28)5(eg)2 that occurs upon reduction must result in a greater degree of bond stretching and possibly stereochemical change compared with Fe(III)/(II) and Ru(III)/(II) couples which involve the conversion (t28)5-+ (t28)6. One of the consequences is presumably the release of "bound" water molecules in going from Co(III) to Co(II). The relatively small As;c (l9 e.u.) found for the Co(sep)3+’/2+ couple (Table 11) could be a consequence of the "tighter" structure of the macrobicyclic ligand effectively exclud- ing specifically bound water in the higher oxidation state. Further- more, the much smaller As;c values for the Ru(III)/(II) and Fe(III)/(II) phenanthroline and bipyridine complexes may arise partly from significant delocalization of the added t28 electron around the aromatic rings. This delocalization can result in additional solvent ordering around these ligands in the divalent state which will counteract the decrease in charge density at the metal center. In the Ru and Fe systems these effects apparently negate each other, yielding As;c values close to zero. 112 It would be enlightening to ascertain the functional form of the ligand effect on As;c, assuming, of course, that such a relation exists. In other words, with respect to the reaction MIIIL'L"+-e- +~ MIIL'L", it is useful to determine how AS° is m n m n rc affected by changes in,m and n. Such information will yield further insight into the various ligand influences on solvent polarization and will allow interpolation and extrapolation of As;c values between related redox couples. The simplest approach to the present question is to consider that the observed value of As;c for a redox couple arises from independent, additive contributions from each ligand in the complex. This partic- ular model also forms the basis of the empirical entropy correlations of George et al. (43,44). The dielectric-continuum Born model could also be used to calculate As;c values, although one can expect some problems since this model treats the nonsymmetrical complexes in this study as uniformly charged spheres. Some of the results of this investigation will be employed to judge the suitability of these simple treatments for predicting the reaction entropies of mixed- ligand complexes (i.e., those containing two different types of ligands). Table 12 summarizes the experimental values of As;c for octahedral mixed ligand redox couples of the type MLéLg(III)/(II) for which values are also available for the corresponding "pure ligand" couples ML;(III)/(II) and ML;(III)/(II). Listed for each couple is the Born reaction entrOpy (as;c) that is calculated with the Born model 113 o mH- om.Hs- -s\-mHzoVoe mn- eN- oaN- -N\-.ao:zoVua N N +N\+M 2%: we m.m m.NH no +N\+MHseHVNaN=oVsm-otoxu m .m m . NH no N 1M HRHVN HNmoVse-nso A.sH on +N\+MHNmoVse m.aH n.sN oN +N\+MHN=oVonszsa-oso m.sH m.HN mN +N\+M=omHmszse or: 3H +NH+MHH=zVsa o.NH m.mH NH +NH+HHoHHHszne as a: +NH+HcoaaaHHszVsa n.0H m.NH m.NH +N\+mxnosHm=zVse o.n N m.o +N\+MHenoVNHmszsa-wsc on H +N 1” Ease o .n.m .nuomnowmqv H .n.o .oumoaowmoV .n.o .m ammo oHonoo xoomm aHHV\AHHHV nHan make on» no moHonoo xoomm mnOHum> How moHnouunm noHuommm .NH oHrme :- 114 Notes to Table 12. aExperimental value of as°rc taken from Tables 9-11 unless otherwise noted. bReaction entropy for mixed-ligand couples estimated from the experi- mental values of the two appropriate pure-ligand couples by linear interpolation (see text). cReaction entropy calculated from the Born model by using equation (8). dThese are estimated values for u a 0.1. Direct determinations at this ionic strength are precluded by hydrolysis (see experimental section). The tabulated values are estimated from those determined in EE‘HpTS (Table 11) by assuming an ionic-strength dependence 3+/ 2+ 2 (Table 11). similar to that of cis-Ru(NH3)2(bpy) eReferences 51, 114, and 121. 115 by using the relation (46): (AS° )Born - 9 65 (22 IE - z2 I- ) (8) , rc ' ox ox red ared where 20x and zred are the net charges and 30x and fired are the equivalent radii (81) of the oxidized and reduced species, respec- tively. (The a value for a complex was taken to be equal to half of the cube root of the product of the diameters along the three L-MrL axes (81). Values of the equivalent radii were taken from reference 81 or estimated from ionic radii (11,18).) Also given for the mixed- eat that were estimated by ligand couples are reaction entropies (As;c) linear interpolation from the experimental values of As;c for the corresponding two pure ligand couples, i.e., by assuming that the two types of ligands provide additive contributions to As;c that are proportional to the number of each ligand type present in the mixed- ligand couples. For the Ru(III)/(II) couples containing ammine and bipyridine ) est ligands, excellent agreement is seen between As;c and (As;c , whereas the Born model fails to predict the marked decreases in A8;c that are observed as the number of bipyridine groups increases, par- 3+/ 2+ 3+/ 2+ 2 3 ° results suggest that differences in local salvation of the individual ticularly in going from c-Ru(NH3)2(bpy) to Ru(bpy) These ligands rather than changes in the average distance of closest approach of the solvent to the metal charge center are largely res- ponsible for the observed large variations in As;c. This finding is perhaps not surprising for complexes containing large ligands such as bipyridine where a number of the solvating water molecules 116 will lie in the vicinity of the aromatic rings away from the metal charge center and the other ligands. As mentioned previously, the addition of an electron to such Ru(III) complexes to form Ru(II) could well give rise to two competing effects upon the surrounding solvent structure. The water molecules close to the ruthenium center, including those surrounding any ammine ligands, will be less polarized and therefore less "ordered" in the lower oxidation state, giving rise to a positive contribution to 15;. However, the water molecules adjacent to the bipyridine rings could well experience an increase in polarization in going to the Ru(II) state since the added t electron will be significantly delocalized around the 28 aromatic rings, acting to increase their net charge density. This latter effect would yield a negative contribution to As;c which will be roughly proportional to the number of pyridine rings. Therefore the stepwise replacement of ammine by bipyridine or pyridine ligands would result in an approximately linear decrease in Aszc, as observed. 3+/ 2+ 3 As;c 3‘’0. The small (or slightly negative) values of As;c observed For Ru(bpy) , these two effects presumably cancel, yielding for Eu(III)/(II), Fe(III)/(II), and Os(III)/(II) polypyridine couples have been previously ascribed to the efficient shielding of the metal center from the solvent by these ligands (51). This last explanation seems less reasonable since the polypyridine ligands will allow some solvent molecules to approach close to the metal center along the Open channels formed by the planar aromatic rings. Also, the markedly larger values of As;c (22 e.u.) observed for both Co(bpy)§+/2+ and 3+ 2+ Co(phen)3 I (Table 11) can be more easily understood on the basis 117 of the present model. These Co(III)'+ (II) reactions involve the electronic conversion t; +t5 e2 which should minimize the extent of 3 233 electron delocalization in the reduced state, therefore discouraging any increase in solvent polarization in the vicinity of the pyridine rings. This change in electronic configuration will also yield a marked expansion of the cobalt center (127) and hence an especially large decrease in the polarization of nearby water molecules. We have obtained further evidence that favors the present inter- pretation from the observation that the ferrocinium/ferrocene couple has a distinctly negative value of as;c in aqueous media {-5 e.u., Table 11). Since this couple is of the charge type +1/0, the simple dielectric-polarization model predicts a small positive value of As;c. The negative As;c can be explained by noting that the electron added in the reduction will be substantially delocalized around the cyclo- pentadienyl rings in a manner similar to the polypyridine couples. Furthermore, the "sandwich" structure of the ferrocinium/ferrocene couple should largely prevent the close approach of solvent molecules to the metal center. Consequently in this case the predominant con- tribution to As;c is anticipated to be the increased polarization of water molecules adjacent to the aromatic rings in the lower oxidation state, in accordance with the observed negative value of As;c. Negative values of As;c have also been observed for some blue copper protein couples having the charge type +l/O and have been interpreted in terms of hydrophobic effects in the vicinity of the copper redox center (128). 118 As noted previously, the large values of As;c typically observed for aquo couples have been attributed to field-assisted hydrogen bonding between the aquo ligands and the surrounding water molecules. This effect is apparent when it is noted that the substitution of a single aquo ligand into Ru(NH3)2+/2+ yields a proportionately larger increase in ASrc than those which result from subsequent aquo substi- tutions (Tables 10 and 12). However, there is reasonable agreement 0 o estd 3+/ 2+ between ASrc and (Asrc) for o Ru(NH3)4(OHZ)2 . 0n the other hand, substitution of one bipyridine in Ru(bpy)§+/z+ by two aquo ligands yields somewhat smaller values of As;c compared to (Assc)estd (Table 12). It is possible that the extent of hydrogen bonding involv- ing aquo ligands is very sensitive to the electrostatic and steric environment. Therefore, it may be generally difficult to achieve accurate predictions of As;c values for mixed-ligand couples containing aquo ligands. Since a large number of complexes employed in redox kinetics contain anionic ligands, it is of interest to examine the stepwise changes in As;c that occur as anions are substituted into the coordina- tion sphere. Substantial and even qualitative changes in As;c are generally expected from the Born model as well as from local salvation effects due to the variations in the overall charges of the complexes. Unfortunately, suitable data are extremely sparse. Table 12 contains data for three Fe(III)/(II) couples containing bipyridine and/or cyanide ligands. Again it is seen that the estimated value /2 for the mixed-ligand couple Fe(CN)4bpy- - is in reasonable agreement 119 with the experimental value of As;c, while the Born model fails to provide adequate estimates of As;c for all three couples. The assumption that the reaction entropies for mixed-ligand couples arise from simple linear additive contributions from each ligand would appear to provide a useful, though approximate, means of estimating As;c under some circumstances. However, often one or both As;c'values for the corresponding pure ligand couples are unavailable. If it is assumed that values of As;c depend chiefly on the ligands and the charge type of the redox couple, the required reaction entropies could be inferred from those for other couples containing the same ligands. George et al. (43,44) have used ionic entrOpy data to obtain empirical parameters for various ligands which allow estimates of As;c to be made for couples containing these ligands. As was discussed earlier, this approach has serious limita- tions since in some cases As;c can be significantly dependent on the nature as well as the charges of the central metal ions. Nevertheless, the method could be employed, albeit with caution, to estimate the changes in As;c resulting from.minor alterations in ligand composition. 2. Correlations between Outer-Sphere Self-Exchange Rates and Reaction Entropies Aside from yielding information on the thermodynamics of ion solvation, reaction entropies also provide insight into some factors that influence redox reactivity. In order to recognize the origin and the utility of the relationship between As;c and homogeneous electron- transfer rates, one must recall that the free energy of activation 120 A631 for an outer-sphere self-exchange reaction consists of inner-shell (AGZn) and outer-shell (Acgut) contributions arising from the struc- tural reorganization of the reactants and the surrounding solvent, respectively, that are required to yield the activated complex for electron transfer. A correlation is expected between Acgut and As;c since both reflect changes in the solvent structure which take place in the course of a redox reaction. The outer-shell term stems from the alterations in solvent polarization that transpire when the acti- vated complex is formed from the reactants (or more precisely, from the precursor complex). The reaction entropy is a measure of the solvent reorganization which occurs around one of the reactants when it is converted to a product. The Francerondon restrictions on electron transfer tend to yield an activated complex having properties intermediate to those of the reactants and products. Thus, in this simple analysis, the entropy change associated with solvent reorganiza- tion in the activation process is related to some fraction of As;c. It must be emphasized that one cannot expect some sort of direct correspondence (such as an equation) between As;c and Acgut’ After all, the overall AS° for a self-exchange reaction is zero since no net reaction occurs. Therefore, As;c is not linked to the electron- transfer rates via an entropic driving-force effect. Furthermore, Acgut refers to the activation of two reactants, while As;c is con- cerned with just one redox center. Fortunately, the outer-sphere reactions considered here are weakly adiabatic, and the absence of strong coupling between the reactants in such processes allows each redox center to be treated independently from a conceptual viewpoint. 121 (It remains to be seen whether the process of bringing two highly charged species together will lead to effects that neither would produce by itself, even if it is assumed that a weak mutual inter- action occurs between these species. This point is discussed in Chapter V.) In addition to the possible correlation between As;c and Acgut’ one could expect some relation between reaction entropies and AGED. As noted previously (e.g., in the discussion of the As;c values for Co(III)/(II) couples), As;c becomes larger as the difference in the effective radii A5 of the reduced and oxidized species increases. Since large values of A; are also associated with large inner-shell contributions to AGfl, some factors that tend to increase AGin are expected to increase As;c as well. Thus, one generally cannot use ASZe to separate the inner-shell and outer-shell components of AGfl, but some correlation between As;c and A631 should be found. The interpretation of AGfl is simplest for redox couples which undergo weakly adiabatic electron transfer and for which the change in metal-ligand bond distances is not large. For such systems the inner-shell contribution AGin is small and calculable, and the outer-sphere contribution A63“ can be obtained from AGf1 (81). t Although suitable systems are not abundant, Ru(III)/(II) couples con- taining ammine and/or polypyridine ligands provide such a class (81). 3+/2+ at NH b Estimates of AG I for various Rn( 3)2x( py)3_x couples where x - 0-3 have recently been made (81). These are listed in Table 13 along with the corresponding values of AG:ut (calcd). The latter 122 .NH oHnoH scum cmxwu .Hoooa anon 0:» Scum wouoHooHoo oHeooo xowou mo haouuao GOHuooom .Az H.o u : cuwdouum 0H:0H um wocHounoV HH was 0H moHnma scum cQHMu .oHesoo xooou mo haouuoo doHuomom .Hw mucouomou mo >H oHAMH scum coxmu “Amnv mucosa moons: scum wouoH=UHmo :OHuoaHuucoo HHoSquouso .Hw mucouowou mo >H oHnmy Bonn ooxou moon .muoouomou wouonmdom onu scum onmEoo oonHHHoo can show 0» wouHoomu Mao; UHuoumouuooHo osu mH MB woo .oOHuanuucoo HHochuoccH onu mH dwoq .ucoumdoo «you co>uomno onu mH HHx .muamuomou mo uHoe HoHooHuuoe .u sH H uoo o3» yam hoooovouw :OHmHHHoo osu mH N muons 3 I «Uq I A H& aH I N dHVHm n auq scum ooonuno .a0Huooou owoozoonMHom o£u Mom hmuodo ovum :OHuo>Huoo ass on :OHuonHuuooo Auso>Homv HHochuouoo .Hw ouomuomou scum :oxou .aOHuoo waHuommu mo musmu unoHo>H=vm 0 o o a m m.cH m.mH m.o m.n m.m oAmmzvsm +~ \+m o o o 0 “mm m a 0 NH AH a n w m m m +N\+m A mzv m o o o o o h“ .V m d.— m 0H m NH H m H m o q +~\+mA Av A mzv m N N m o o o o o h“ a o m n o o q o q o m +N\+h A av A mzv m m o o o s %Q H.— c A H m m o m m o +N\+m A Av m .:.o.7 .s.o .Hoa.Hoox .HoE.Hoox < oHaoou xooom HI HI o AoOHoovoum< .oumq .AooHoovuoou< .uooo< .o \:0Huooom owdmnoxm o e w e o a A a o I .oomu um ooHasoo xoowx wdHocoemouuoo on» new moHdouuom :OHuooom onu woo meHuooom omdo£OXMIMHom AHHv\AHHHV ESHGonuom meow you monuocm moum :OHuo>Huo< dooSqu oowwuoefioo .MH oHooH 123 quantities were calculated by using the Marcus model which treats the surrounding solvent as a dielectric continuum (81). The values of As;c for these ruthenium couples are given in Table 13 along with the corresponding values of A8;c(calcd) calculated using the dielectric-continuum Born model (see reference 81 and equation (8) from the previous section). Table 13 also includes values of a, the equivalent radius of the reacting cation. Because of the very similar a values found for the trivalent and divalent Ru complexes considered here (81), one can list a single value for both halves of the redox couple. It is seen that the increasing values of A63“t that are observed as the smaller ammonia ligands replace bipyridine in the ruthenium coordination sphere are also associated with markedly increasing values of Asgc. This is not surprising since greater changes in solvent polarization are required for electron transfer as the effec- tive radii of the reactants decrease (79,81). The dielectric- continuum estimates of AGgut(calcd) and As;c(calcd) are seen to be in broad agreement with the corresponding experimental parameters. However, it is interesting to note that there is substantially better agreement between Acgut and AGgut(calcd) than between As;c and As;c(calcd). This behavior may be related to the fact that Acgut(calcd) is mainly determined by the optical dielectric constant 80p (see reference 79 and equation (12) from Chapter I), rather than by the static dielectric constant as. The latter determines the value of As;c(ca1cd) (46). It is known that eop is much less sensitive to 124 solvent structure than is as, so that any breakdown in the validity of the dielectric-continuum model of ion-solvent interactions caused by an alteration in the solvent structure in the vicinity of the solute is likely to affect acgut to a much smaller extent than As;c. Most other outer-sphere exchange reactions involve significant changes in.metal—ligand bond distances so that the inner-shell reorganization energy AG:n forms a sizable component of AGfl. This is particularly true of M(OH2)2+/2+ reactions, and some reactions of this type are listed in Table 14. This table also includes values of the self-exchange rate constant kll’ AGfl, and As;c for the aquo couples and a number of other 3+/2+ redox couples containing ammine, ethylenediamine, or bipyridine ligands. (The tabulated values of the activation free energy were corrected for the electrostatic work wr required to form the collision complex by using the equation AGEl - RT(ln Z - 1n kll) - wt where Z is the bimolecular collision frequency (81)). The.ammine and ethylenediamine complexes, in par- ticular, form an interesting comparison with the aquo complexes since all three ligands are of comparable size so that the outer-shell terms Acgut should be similar, at least on the basis of the dielectric- continuum treatment of Marcus (79). The data from Tables 13 and 14 are also presented in Figure 4 as a plot of As;c versus AGfl. Inspection of Table 14 and Figure 4 reveals that there is a general tendency of Asgc to increase along with AGll’ although not surprisingly there is a considerable scatter in the individual points. For the aquo couples, the increasing values of As;c as well as of AG:1 are 125 Table 14. Comparison between Kinetic Parameters for Selected Outer-Sphere Self-Exchange Reactions and Reaction Entropies for Corresponding Redox Couples at 25°C. k a * b c Exchange Reaction/ ex A611 As;c Redox Couple 'Muls-1 kcal.mol: e.u. «(01196343 ~ 2 x 10'7(1.0)133 ~ 24 so V(OH2)63+/2+ 1.5 x 10‘2(2.0)13S 17.4 37 Fe(OHz) 63442“ 4(0. 53136 13.8 43 Ru(OH2)63+/2+ ~ 80(1.0)132 ~ 12 36 Ru(NH3)63+/2+ 3 x 103(o.l)81 8.2 18.5 Ru(en)33+/2+ 3 x 103(o.1)81 8.9 13 1111(1)”):+ I 2+ 4 /x 108(o.1)137 ' 3.4 1 Co(en)33+/2+ 8 x 10"5(1.0)138 20.4 38 Co(sep)3+/2+ 5(o.2)88 14.4 19 Co(bpy) 33”“ ~ 20(o.1)139 13.4 22 8Rate constant for acid-independent self-exchange pathway; obtained from indicated reference numbers. Ionic strengths are given in parentheses. bFree energy of activation for the self-exChange reaction, corrected for the electrostatic work required to form the collision complex. cReaction entropy of redox couple, obtained at (or extrapolated to) ionic strength u I 0.1 M. Data from Tables 9-11. 126 .+N\+m Assc A secee AmAV .+s\+mAAesceAMszvse ANAV .+N\+WAN Nmoveo AoAV .+N\+mAN mov> Aav .+N\+mMN moose Amy .+N\+WAN eaves AAV .+N\+mAsseVoo Ase Assess Ame .+~\+MAAeevos Ase .+N\+eAAszces Ame mAssess ANV .+N\+mAAesvsa AAV .esoseAs on Ass .+N\+m :AAsevsA secse AAAV .+N \+m .+~\+n .moHeooo xooou or» you maouuco oOHuooou can .o> mooHuocou owdmcoonwHom ouoaemluouoo osom How mwuooo cOHum>Huoo osu mo uOHm .o oustm 127 Om .c oustm :Les Toss .818“? 0m 08 on ON 0. _ _ _ _ _ O. O . m _ ON mm HED<7 (._Iow loom 128 broadly compatible with the conventional explanation which ascribes the latter changes to increasing values of the inner-shell contribu- tion AG* . Thus the Cr(OH2)3+/2+ exchange reaction involves transfer in of an antibonding e8 electron which is expected to yield substantially 3+/ 2+ 3+]2+ larger values of A; than the Ru(OH , Fe(OH and 3+/ 2+ 2)6 2)6 exchange reactions which involve transfer of a nonbonding 3+/ 2+ V(OHZ)6 electron. Indeed, the Cr(OH reaction exhibits the largest t23 2)6 values of both A611 and ASrc . However, substantial differences in A611’ which do not correlate in any simple way with As;c, are observed between the Ru(OH2)3+/2+, Fe(oaz)3+/2+ and V(OHZ)3+/2+ reactions. Of these aquo couples, A; has been directly determined 3+/ 2+ using crystallographic data only for Fe(OH2)6 ; a value of 0.14A was obtained (129). This value of As leads to a calculated value of k11 that is in good agreement with the measured value for the 3+/ 2+ 3+/ 2+ Fe(OH system (130). The approximation of A5 for V(OH 2)6 2)6 can be obtained by noting that the difference in ionic radii for six- coordinate 173+”+ is 0.15 A (11). The corresponding difference for Fe3+lz+ is 0.14 A (11) which agrees well with A5 for Fe(OH 3+]2+. 3+/ 2+ 2)6 Thus, A; for V(OH is expected to be slightly larger than that 3+/ 2+ 2)6 for Fe(OH2)6 , but this does not completely account for the sub- stantially slower exchange rate of the vanadium.couple. Furthermore, the A; values for these two couples would lead one to predict that V(OH2)3+/2+ have a larger AS° cthan Fe(OH2 )3.”2+ which is the opposite of what is actually found. One factor which complicates the interpretation of the results 129 for the aquo redox couples is the lack of information concerning the coordination numbers of many of the ions of interest. Thus, as noted previously, the vanadium results can be explained by an increase in the number of bound water molecules which could well accompany the reduction of V(III) to V(II). This increase in the solvent struc- turing around the V(II) ion will result in a smaller value of A8;c, as is observed, and will yield an additional contribution to AGE1 which will act to decrease the self-exchange rate. Similar considera- tions can be applied to Cr(II). The aquo ligands in the axial positions of the hydrated ion are weakly held as a consequence of Jahn-Teller distortion (as mentioned in the previous section). Thus, the reduction of Cr(III) to Cr(II) will result in an effective decrease in the number of bound water molecules. This change in the coordination shell could be responsible in part for the large As;c and 3+/ 2+ 6 O The data in Table 14 and Figure 4 reveal-that-a dependence of A611 Ac;1 for Cr(OHz) upon A3;c is also obtained for non—aqua redox couples. Thus both AGI1 and As;c are markedly larger for Co(en)§+/2+ than for Ru(en)§+/2+; similar differences are also seen between Co(bpy)3+/2+ 3 3+]2+ . and Ru(bpy)3 . It is likely that the larger values of ASrc for the cobalt couples are related to their larger A5 values. (As an example of the size of the values which are involved, As - 0.18 A for Co(NH3):;+/2+ while the corresponding quantity for Ru(NH3):g+/2+ is 0.04 X (127).) The larger values of Acgn and hence of A611 for the Co systems are at least partly a result of these larger A5 values (131). 130 The observed differences can also be discussed in terms of electronic structure. The cobalt(III) complexes are low tgg spin while the cobalt(II) complexes are high spin tgg e2, so a spin multi- plicity change accompanies the electron transfer. The ruthenium(III) and ruthenium(II) complexes are all low spin, and as a result no spin multiplicity change accompanies the electron transfer. A consequence of the spin multiplicity change in the cobalt system is nonadiabati- city (that is, K, the probability of electron transfer within the activated complex, is < 1; nonadiabaticity increases the apparent value of AG* since AG*1 can be considered to include the term 11 1 -RT in K). It has been estimated (131) that K for the Co(NH3)3+/2+ 6 exchange reaction is ~v10.4. If a similar factor applies to the 3+I2+ exchange reaction, AG;1 for the ethylenediamine couple should be decreased by about 5 kcal mol-l. Although this correction Co(en) is appreciable (and reduces the scatter in Figure 4) substantial dif- ferences between the cobalt and ruthenium systems remain even after corrections for spin multiplicity. Obviously other factors are also Operant. The different ligand conformations in Co(en);+ and Ru(en)§+ (126) would have an effect on the inner-shell reorganization of these complexes and hence on A031. This difference could also be partly responsible for the observed A8;c values since the availability of the metal ion to the solvent will be affected by the conformations of the ligands around the cation. Some of the variations between Ru(bpy):;+'/2+ and Co(bpy)?”2+ can be explained in terms of the delocalization of the added electron in Ru(bpy)§+ around the aromatic rings. As discussed previously, the 131 2+ 3 this complex. In Ru(bpy)§+, the increased charge density of the electronic structure of Co(bpy) precludes a similar effect within ligands induces an increase in the degree of structuring in the solvent immediately adjacent to the aromatic rings. Thus the reduc- 3+/ 2+ 3 solvent ordering, and this provides some explanation for the small tion of Ru(bPY) to Ru(bpy)§+ is accompanied by a small change in As;c and A631 values of this couple. One can conclude that effects other than A; and nonadiabaticity give rise to many of the differences between the Co(III)/(II) and Ru(III)/(II) systems. Unfortunately, it is more difficult to treat the ligand-conformation and electron— delocalization factors in a quantitative manner. The differences in As;c for Fe(OH2)2+/2+, Ru(OH2)g+/2+, and Ru(NH3):+/2+ have recently been used to provide an estimate of ) 3+/ 2+ 2 6 depends upon the extent to which As;c for the aquo couples differs Aaa~0.l R for Ru(OH (132). The accuracy of this estimate from similar ammine couples due to the hydrogen-bonding ability of ) 3+] 2+ 2 6 the empirical entropy correlation of Powell and Latimer (41) predicts the aquo ligands. Taking A; for the Fe(OH couple as 0.14 A, that A8;c should only be about 10 e.u. lower for an analogous aquo couple for which A5 - 0. However, A8;c for the Ru(NH3):;+/2+ couple (A; ' 0°04 3 (127)) is 25 e.u. smaller than for Fe(OH2)2+/2+ (Table 14). Most likely, therefore, hydrogen—bonding effects do con- tribute to As;c for the aquo couples. Despite the complexities, there appears to be a consistent cor— relation between AGfl and Asrc for the outer-sphere exchange reactions 132 considered here. The apparent simplicity of this correlation is somewhat deceiving since some of the factors which contribute to A8;c may not always affect AG:1 in a similar way. Nevertheless, the cor- relation could be used in carefully selected systems as a means of estimating A; and hence Asia. More importantly, the relationship supplies further clues to the nature of the solvent-rearrangement process required for electron transfer. CHAPTER V ACTIVATION PARAMETERS FOR HOMOGENEOUS OUTER-SPHERE ELECTRON-TRANSFER REACTIONS: COMPARISONS BETWEEN SELF-EXCHANGE AND CROSS REACTIONS USING THE MARCUS THEORY 133 A. Introduction As noted in Chapter I, the adiabatic model of outer-sphere electron transfer developed by Marcus and others predicts that there should be a simple relationship between the kinetics of homogeneous cross reactions and the corresponding self-exchange processes. This relationship (the "Harcus cross-relation") has commonly been formu- lated as equation (1): 1 / k12 ' (kllkzlezflz) 2 (1) log £12 - (logK12)2/4 log (kllkzzlzz) ' (1a) where k11 and k22 are the rate constants for the two constituent self- exchange (homonuclear) reactions, k12 and K12 are the rate and equi- librium constants, respectively, for the corresponding cross (hetero- nuclear) reaction, and Z is the bimolecular collision frequency in solution. Nomerous tests of the applicability of eqn (1) to experi- mental kinetic data have been made. It has frequently been observed that the predictions of eqn (1) agree reasonably well with experimental results (60,140). The observed values of klz, Rigs, are often within an order of magnitude or so of the values, kiglc, that are calculated from k11 and kfiz using eqn (1). However, it has recently become clear that there are a disturbingly large number of reactions for which kggs calc and klz are in substantial disagreement (132,134,141—145). For most of these systems, it is found that Rigs <3 cgleand that the difference between k;:8 and kiglc increases as the equilibrium.constant K12 increases (134,142-144). 134 135 Another puzzling feature of homogeneous outer-sphere redox processes between like-charged ions is that the entropies of activation AS* for self-exchange as well as cross reactions are typically large and negative (ca. -20 to -35 e.u.) and insensitive to variations in ionic strength (81,144b). Such values of A8* are predicted by simple electrostatic theory at low ionic strengths, but I As for self-exchange reactions is theoretically predicted to increase markedly with increasing ionic strengths to a value close to ~10 e.u. (60,81,146). In view of these striking discrepancies between theory and experi- ment, it seems worthwhile to examine the capability of the Marcus model to predict the relative values of the free energies, enthalpies, and entropies of activation for corresponding self-exchange and cross reactions. Such an investigation should help to pinpoint the factors that are responsible for the observed breakdown of eqn (1). Few tests have been made of the ability of the Marcus model to correlate activa- tion parameters for corresponding self-exchange and cross reactions (134,145) because many of the required enthalpies and entropies of the cross reactions, AHiz-and A812, have been unavailable. However,the standard (formal) potentials and reaction entropies tabulated in Chapter IV can be used to obtain accurate AH;2 and A832 values at the ionic strengths that are typically employed for kinetic measurements. In the present work, these thermodynamic quantities are utilized to compare the experimental activation parameters for a range of outer- sphere cross reactions with those that are predicted by the Marcus 136 treatment from the kinetic parameters of the corresponding self- exchange reactions. This analysis will provide an additional insight into the limita- tions of the Marcus and related models in describing the energetics of homogeneous outer-sphere electron-transfer processes. B. Kinetic Formulations and Results 1. Free Energies of Activation For the present purposes, eqn (1) can be usefully rewritten in terms of free energies of activation (134,147): * a * * ° A612 0.5(AG11 + A622) + 0.5(1 + a) A612 (2) 3 ° * * a A612/4(A611 + AGZ (2a) 2) where AG;2 and A032 are the "standard" free energy driving force and the activation free energy, respectively, for the cross reaction (determined at the appropriate ionic strength), and AGfl'and A632 are the free energies of activation for the self-exchange reactions. The activation free energies appearing in eqn (2) should be corrected for the work of forming the collision (precursor) complex from the separa- ted reactants since they are actually reorganization energies within such a binuclear assembly (79). Eqn (2) should therefore be written for the experimentally accessible ("apparent") free energies of acti- vation (AG*)app in the form (79): w w w w * a: * * .. _ (8612)app 0.5[(AGll)app + (AG22)app] + 0.5(AGlz+A621 6611 A622) (6622 + Acgl - Ac¥2)2 + O.SAG{2 + w w (3) * * - _ 8““19st + (AG22)app AG11 A“22] 137 The apparent free energies of activation are related to the experimental second-order rate constants k°bs'by(74.79) ksbs . z expl-(AG*)app/RT] (4) where Z is the bimolecular collision frequency. In eqn (3), A312 and ACgl are the free energies required to form the precursor and successor ("collision") complexes from the separated reactants and products, respectively, and A621 and ACgZ are the corresponding work terms for the constituent self-exchange reactions. For cross reactions between like-charged ions of similar structure, it is expected that W ‘W W W W A012 ~ A021 ‘3 A611 *3 A622 - AG , so that eqn (3) can be simplified to s . * + * + ° (A012) p 0.5[(AG11) p A022) p ] 0.5 (l+a)AG12 (5) a o I _ W In principle, the presence of positive values of ACW in eqn (5) can explain the common observation that cation-cation reactions with large driving forces have observed cross-reaction rate constants kggs are calc substantially smaller than the values k12 that are calculated using eqn (1). For small values of Asz, force term in eqn (5) will generally be small since aAsz is propor- the influence of a on the driving- tional to the square of Asz (eqn (5a)), so that eqn (5) will approxi- mately reduce to eqn (2) under these conditions, irrespective of the value of ACW. However, as the magnitude AG:2 increases, the effect of 6 upon the driving-force term will increase so that the presence of positive values of AC” will act to enlarge progressively the values 138 obs calc of (A612)app calculated from eqn (5). This yields k12 <3k12 since eqn (1) and its free energy analogue, eqn (2), do not include work terms. The above argument is qualitatively reasonable, but the usual procedure of calculating work terms from the Debye-Hflckel model is inadequate in a quantitative sense because the values obtained, (ACW)DH, are not large enough to explain the extent of the observed discrepancies (134). Nevertheless, it seems quite plausible that another, larger component of AC", ACE, could arise from the mutual solvent ordering which accompanies convergence of the two cationic reactants. Consequently eqns (5) and (5a) can conveniently be re- written in terms of "Debye-Hflckel corrected" free energies of activa- * 0 tion (AG )corr' * u s * ° (AG12)corr 0.5[(Acn)corr + (AG22)corr] + 0.5(1+G)AG12 (6) ‘W a: o * _ a AG12/4[(Acll)corr + (A622)corr ZAGs] (6a) where (AG*) - (AG*) - (AGW) - AG* + AGW (7) corr app . DH 5 w lezezN and (81) (AG ) - T (8) DH 0 ° Ib . . esa(l+Bau ) where 21 and 22 are the charges of the two reactants, e is the elec- tronic charge, N is Avogadro's number, as is the (static) dielectric constant, B is the Debye-Hfickel parameter (82), u is the ionic strength, and 3 is the distance between the centers of the reacting ions in the collision complex. Eqns (6) and (6a) will be employed as a convenient means of examin- ing the relationship between the experimental activation free energies 139 of corresponding self-exchange and cross reactions in terms of the Marcus model. These equations also provide the basis for exploring the behavior of the individual enthalpic and entropic components. Although the hypothesis embodied in the use of the term A0: in eqn (6a) cannot be proven, it will be shown that the algebraic forms of eqns (6) and (6a) are nicely consistent with.most of the available experimental data. Tables 15 and 16 summarize the rate parameters for the acid- independent pathways of a number of homogeneous self-exchange reactions between various cationic complexes. Table 15 also includes Ef and A8;c values for the redox couples involved in these reactions. The (or AG* ) were obtained in the followb * listed values of (A51 22)corr l)corr ing ways. For reactions for which the self-exchange rate constants kg?“ have been determined, values of (AG* 11)corr were obtained by using eqns (4), (7), and (8), assuming that Z - 6 x 1010M.1 sec.1 (152,153). obs 3+/2+ 3+]2+ 3+]2+ The listed values of kll for Ruaq , Ru(NHB)6 3 , 3+/2+ self-exchange are quoted from the literature (81, 1329 , Ru(en) and Ru(NH3)5pY 145) and_were obtained from cross-reaction rate constants by using eqn (1). These cross reactions had small driving forces and involved the Ru complex of interest and a structurally similar coreactant. The justification for this method of calculation is that eqn (1) has been found to be consistently successful under these conditions (134). values of the "Debye-Hfickel corrected" rate constants kiirr are also given in Table 15. These were obtained from kiis using the relation corr obs W ob lnk11 . lnk11 + (AG )DH/RT. No quantitative values of klls are 140 Table 15. Kinetic and Thermodynamic Parameters of Some Self-Exchange Reactions at 25°C. a b obs c corr Er ASre k11 k11 Redox _1 _1 _1 _1 Couple mV. vs. a.c.e. e.u. M. sec. M. sec. Caz/2+ 1680148 ~45e 3. 3(3) 151 8 Pei/2+ 500(o.2) 43 4(0.5)136 15 Ruiz/2+ -15(0. 3) 36 ~60(l)132 200 viz/2+ -475(o.2) 37 0.015(2)135 0.03 Buzz/2+ -625(0.l) 48 (4 x 10“) Griz/2+ -660(1) 49 (2 x 10'6) fizz/2+ -1425(o.l) 48 <~o.l) 114""3+ -880(0.5) 48 (0.5) as ling/3+ -80(1)149 52150 (~o. 02) Ru(NH3)2+/2+ -180(0.2) 18 ~3 x 103(o.1)81 ~s x 10“ Ru(en)?/2+ -60(0.1) . 13 ~3 x 103(o.1)81 ~5 x 10" Co(en)§+/2+ ~460(l) . 37 8 x 10'5(1)138 2.5 x10‘4 Co(phen)?/2+ 145(0.05) 22 ~40(o.1)139 ~150 Co(bpy)?/2+ 70(0.05) 22 ~ 20(o.1)139 ~80 Ru(bpy)?/2+ 1045(o.1) 1 2x 109(1)137 ~1 x 1010 Ru(N33)5py3+/2+ 75(0.l) 17 4.7 x 105(1)MS 1.5 x 106 141 Notes to Table 15. aFormal potential of redox couple in mV vs. KSCE. Ionic strength u given in parentheses. Data from Tables 9-ll except where otherwise noted. For most systems, E becomes only ~5 - 10 mV more negative f ‘with increasing u over the range u “v 0.1 to 1.0. Superscripts are the reference numbers of the data sources. bReaction entropy of redox couple. Data from Tables 9-ll except where otherwise noted. Quoted values were obtained at the same ionic strengths as E Superscripts are the reference numbers of f. the data sources. cObserved second-order rate constant for the acid-independent pathway. Corresponding ionic strength given in parentheses. Superscripts are the reference numbers of the data sources. dRate constant corrected for the Debye—Hfickel work (AGw)DH of forming the collision complex from the separated reactants (eqn (8)). Quoted corr obs W 9 values are obtained from 1n k11 ln k11 + (AG )DH/RT. The a values used in eqn (8) were estimated from the sum of the radii of the reactants. Estimated rate constants are given in parentheses (see text). eEstimated value. 142 Ru(NH3)Spy Table 16. Kinetic Parameters of Some Self-Exchange Reactions at 25°C. a b c d corr * * * kll (AGll)corr (AHll)corr -(Asll)corr - Redox Couple Mtlsectl kcal.mol:1 kcal.mol-.-1 e.u. c63+"2+ 8 13.5 10.2 11 as . Fe3+l2+ 15 13.1 8.6 15 aq Ru3+l2+ 200 11.6 ..7.08 —~153 aq v34”+ 0.03 16.8 ~12.5 15 89 Buzz/2+ (4 x 10'4) 19.3e ~15.og .1158 Griz/2+ (2 x 10’6) 22.56 ..18.03 ~15g szzlz+ («v0.1) ~15.5e ~11g '2153 U4+/3+ (0.5) 15.1e A.9.o .~2oh aq Npgz/3+‘ (1.0.02) «4.17.0f evllh .Nzoh Ru(NH3)g+/2+ 7~5 x 104 8.3 3.2 17 Ru(en)3+/2+ ~5 x 104 8.3 3.2 17 Co(en)g+/2+ 2.5 x 10"4 19.6 13.4 22 Co(phen)§+/2+ ~.1so 11.7 4.5 24 Co(bpy)§+/2+ e-8o 12.1 7.0 17 Ru(bpy)§+/z+ tel x 1010 ~1 A.li .VOi 3+/2+ 1.5 x 106 6.3 2.4 13 143 Notes to Table 16. 8Rate constant corrected for (ACW)DH. values taken from Table 15. For further details see notes to Table 15. bFree energy of activation corrected for Debye-Hfickel work term. values calculated from kiirr by using eqn (4) with Z a’6 X 1010M 1s 1 (152). (Eqn (4) expresses a relationship between k0 ”b and (AG*) p’ but the equation can be used with equal validity for kF°rr and (AG*) .) corr cEnthalpy of activation corrected for Debye—Huckel work term. values obtained from experimental activation enthalpies by using eqn (13) ‘W and (AH11)corr-(AH11)aPp-(AG )DH (see text). Except where otherwise noted, the experimental enthalpies were obtained from the same data source as used for the corresponding kiisin Table 15. dEntropy of activation, obtained from experimental activation entropies 68* by adding 10 e.u. (eqn (12)). Experimental values are from the same data sources as the corresponding kiis values listed in Table 15 (unless otherwise noted). eEstimated from electrochemical exchange rate data (see text). fEstimated from the kinetics of the Ru(en)§+ - szz reaction (Table 19) by using eqn (7) (see text). 8(AS *1) assumed to be -15 e.u. (AH was calculated from corr ll)corr (AG ll)corr on the basis of this assumption (see text). h (Asll)corr assumed to be -20 e.u. (AH was calculated from ll)corr (AG* 1corr1) on the basis of this assumption (see text). 1Estimated value (134). 144 apparently available for the aquo couples, Buzz/2+, Griz/2+ .Yb3z/2+ 4+l3+ 4+/3+ an fld Np aq . However, self-consistent estimates of (A611)corr for these reactions were obtained as follows. The electrochemical exchange rates for these and other aquo redox couples have been deter- mined at the mercury-aqueous interface and have been found to yield a consistent correlation with the rates of corresponding homogeneous self-exchange and cross reactions (94,154). This can be expressed as (AG*):°rr - o.47(6c;*):‘or - o. 5 kcal (94), where (AG*):°rra and (AG*)2orr are the corresponding electrochemical and homogeneous free energies of activation, respectively, that have been corrected for the apprOpriate electrostatic work terms. The listed values of (AG 3+/ 2+ Cr3+/2+ Yb3+]2+ and U4+/3+ aq aq aq 89 A justification for employing this method is that the resulting esti- ll)corr for Eua were obtained in this manner. mates of (AC are consistently within ca. 0.5 kcal of the values ll)corr estimated (by use of eqn (6)) from the kinetics of cross reactions having small driving forces (see below). The listed value of (Acll)corr for Np2+l3+ was determined from the kinetics of the Ru(en)3+ 'NP3: reaction (Tables 19 and 20) by using the latter method, since the electrochemical exchange rate for this couple is not available. Dis- cussion of the activation enthalpies and entrOpies will be reserved for a more appropriate time (i.e., in section 2). Tables 17 and 18 list the available kinetic parameters for the acid-independent pathways of a number of cross reactions between the various aquo couples listed in Tables 15 and 16. values of the "Debye-Hflckel corrected" free energies of activation (AG were 12)corr 145 Table 17. Between Aquo Cations at 25°C. Driving Forces and Rate Constants for Cross Reactions C C a i 6:38 his" Oxidant Reductant kcal mol-1 Mulsec"1 MZ-lsec-1 C03+ Fe2+ 27.2 50(1)155 150 c63+ v2+ 49.5 9 x 105(3)156 2 x 106 Cb3+ Cr2+ 53.9 1.3 3 104(3)156 2 x 104 C03+ 03+ 58.9 1.1 x 106(2)1448 3 x 106 Fe3+ Ru2+ 11.9 2.3 x 103(1)132 7 x 103 Fe3+ v2+ 22.5 1.8 x 104(1)144b 6 x 104 Fe3+ Eu2+ 25.9 7 x 103(1)”7 2 x 104 Fe3+ Cr2+ 26.7 2.3 x 103(1)158 7 x 103 Fe3+ U3+ 31.7 4 x 105(2)1448 8 x 105 Ru3+' v2+ 10.6 2.8 x 102(1)132 9 x 102 Np4+ v2+ 9.1 1.3(1)159 4 v3+ Eu2+ 3.5 9 x 10‘3(2)16o 2 x 10' v3+ U3+’ 9.2 85(2)161 250 Eu3+ Cr2+ 0.8 .12 x 10‘5(0.5)160 8 x 10‘ cr3+ 03+ 5.0 6.2 x 10'2(2)144° 0.15 146 Notes to Table 17. 3Free energy driving force for cross reaction, determined from the formal potentials listed in Table 15 by using AG;2 3 IKE:x - EEed) where Box and Ered are the formal potentials of the couples under- f f going oxidation and reduction, respectively. bObserved second-order rate constants for acid-independent pathway. Corresponding ionic strength given in parentheses. Reference numbers of the data sources appear as superscripts. cRate constant corrected for the Debye-Hflckel work of forming the collision complex from the separated reactants. See note (d) to Table 15. 147 n.0I H.o w.nH o.m +m= +muu N.OI o n.0NII w.o +~uo +m=m o m.o e.HH N.o +mD +m> A.o o o.~H m.m +N=m +m> A.H m.o m.nH H.a +~> +~ez m.H m.o A.OH c.0H +N> +mnm m.o m.w m.e A.o A.Hm +m: +mom ¢.m H.m m.~ m.a A.o~ +~uo +mom w.» o.m o.~ w.w ¢.m~ +Nom +mom N.w m.c H.N N.w m.NN +~> +mom H.m 5.0 m.a a.HH +~=m +mom H.¢ H.HN N.mH ¢.m a.wm +m= +moo 0.5 o.AH H.0H 0.x a.mm +~uo +moo m.s e.mH H.oH H.o m.me +N> +mou N.¢H e.uH m.m A.HH ~.AN +~om +moo HIHoa Hoax HIHoE Hoax HIHoa Hoax HIHoE Hoax HIHoa Hoax odouoswom udowao u “be wNmusen.o u Nme<.em.o neweuANmosv stusu .oenN us o:0Huoo osu< sooauom odOHuuoum ooouu now noouom de>Hua was doHuc>Huo< mo monuocm ovum .mH oHooe 148 Notes to Table 18. 3Free energy driving force for cross reaction, taken from Table 17. bFree energy of activation, calculated from kigrr .using eqn (4). See note (b) to Table 16. (Table 17) by coo determined from.eqn (6a) by using the (AG* values for the ll)corr W apprOpriate self-exchange reactions (Table 16) and ACS - 0. (Table 16 contains (Acll)corr values but these can be used as * * n II II II (AG 11)corr or (A022)corr in eqn (6a) since the 11 and 22 subscripts only serve to distinguish one of the constituent self- exchange reactions from the other.) dCalculated from eqn (6) by inserting the appropriate values of * * * (AG ll)corr and (A622)corr (Table 16), and (A612)corr’ and AGi2 (this table). eApparent work required to form the collision complex from the separated reactants; determined from the listed so and a values by using eqn (9) (see text). 149 were again obtained from.the experimental rate constant kiss by using eqns (4), (7), and (8). Tables 19 and 20 summarize the corresponding data for cross reactions involving the non-aqua redox couples that are listed in Tables 15 and 16. Also included in Tables 17-20 are the free energies of reaction AG;2 for the cross reactions. These latter quantities were determined from the difference in the formal potentials E at 25°C that are given in Table 15 for the appropriate f pairs of redox couples. These values of Ef were obtained by means of cyclic voltammetry (see Chapter IV) using suitably non-complexing media at ionic strengths that are comparable to those employed for the corresponding kinetic measurements. [Although some values of Ef were determined at ionic strengths different from those employed for the kinetic studies, this is of little consequence since the variation of Ef with u is typically small (f;10 mV) in the range of ionic strengths (u5'0.l-l) used for the kinetic measurements. Furthermore, these differences will tend to cancel when the pair of Ef values used to calculate A612 are determined at similar ionic strengths.] The experimental values of (A612)corr for each cross reaction given in Tables 18 and 20 were then compared with the predictions of eqns (6) and (6a) using the following approach. The quantity 0.5010;2 was calculated for each cross reaction by inserting into eqn (6) the (AG AG° * ll)corr’ (A622)corr’ and 12 are listed in Tables 16, 18, and 20. An analogous quantity, which appropriate values of (Asz) that corr’ o * . will be designated 0.5oOA612, was determined from (Acll)corr’ (AG* , and AG;2 by using eqn (6a) with the assumption that 22)corr 150 Table 19. Driving Forces and Rate Constants for Cross Reactions Involving Nan-Aqua Cations at 25°C. a obs b . corr c ‘Aclz k12 1‘12 Oxidant Reductant kcal mol-1 M-J'sec_1 Mulsec"l Ru(NH3):+ vi: 6.8 1.5 x 103(0.48)162 ~2 x 10" Ru(NH3)2+ 3.1:: 10.3 2.3 x 103(1)163 7 x 103 Ru(NH3):+ or: 11.1 2 x 102(0.2)16“ 2 x 103 1:160:33)?“ vb: 28.7 4.5 x 107(1)163 1.5 x 108 Ru(NH3) 2+ Np2: -2 .3 0.3(1)119 ~ 2 Ru(NH3)2+ 0:: 16.0 1.5 x 105(1)165 9 x 105 Ru(en)§+ NPZ: 0.5 8.5(1)119 40 Ru(en)§+ (I: 18.8 8.4 x 105(1)165 4 x 106 Ru(NH3)5py3+ a: 12.7 3 x 105ml“ 1 x 106 Ru(NH3)5py3+ Hui: 16.1 5.4 x 104(1)134 1.5 X 105 Ru(NH3)5py3+ Ru(NH3)§+ 5.9 1.4 x 106mm“ 4 x 106 Fez: Ru(NH3)§+ 15.7 3.5 * 105(o.1)166 ~6 x 106 R1122 Ru(NH3)§+ 3.8 1.4 x 104(1)132 4 x 10" Fez: Ru(en)§+ 12.9 1.4 x 105(0.1)166 ~3 x 106 Fez: M(NH3)5py2+ 9.8 5.8 x 104(1)”5 1.7 x 105 Co(phen)§+ Rummy? 7.5 1.5 x 10“(0.l)13" ~1 x 105 Co(bpy)g+ “(M3)? 5.8 1.1 x 10"(0.1)134 ~1 x 105 Co(phen)§+ Ru(NH3)5py2+ 1.6 2 x 103mm“ 6 x 103 Co(en);+ a: 0.3 7 x 10“'(1)167 2.5 x 10‘3 Co(en)3+ Eu: 3.8 4.5 x 10"3(1)168 2 x 10"2 151 Table 19. (cont.) ' a obs b corr c ‘A612 1‘12 k12 -1 -1 -1 -1 -1 Oxidant Reductant kcal mol M sec M sec Co(en)§+ Griz 4.6 3 X 104(1)167 1 X 10-3 06(en)§+ a: 9.6 0.13(0.2)169 2.5 Co(en)§+ 1b: 22.2 4.5 x 102(0.18)170 ~1 x 10“ Co(phen)§+ a: 14.3 4 x 103ml“ ~1 x 10“ Co(bpy)§+ a: 12.6 1.1 x 103(2)171 2 x lo3 bumpy);+ Pei: 12.4 7.2 x 105(1)172 2 x 106 aFree energy driving force for cross reaction, determined from the E ox red values listed in Table 15 by using AG° - F(E - E to Table 17. 12 f f f See note (a) bObserved second-order rate constants for acid-independent pathway. Corresponding ionic strength given in parentheses. denote the reference numbers of the data sources. Superscripts cRate constant corrected for the Debye—Hfickel work of forming the collision complex from the separated reactants. Table 15. See note (d) to 152 m.o N.o e.m m.m +wAmmzvam wmsm m.m N.N e.A m.m A.nA +wAmmzvaa wmue .3 me An an +..N.Ammzvsa +maamAmmzvam 6.6 m.N m.H b.A H.6A ”was +maamAm=zvsa m.A m.o n.e A.NA WM> +mAanAmmzvam on m m.m e.m N.A A.m m.mA +m= +mAsuVam H.o o m.NN m.o ”maz +MAsuvaa an TN .3 es 0...: ”a +mAmmzvaa o o m.eA m.NI Wmez +MAm=zvam e.m o.s «.8 m.m A.mN Wms» +mAmmzvsa m.o m.o N.¢A N.AA ”was +mAmmzvaa as To ma 3: was +MANAA53A a.c- 8.0 as An... ”MA, Mummies HIHos Hoax HIHoE Hoax HIHoE Hoax HIHoa Houx HIHoE Hoax unmanned“ udoono u “be eNmosem.o uNmu<.um.o swuouANmuav u Nmusu .oomN on usofiuuo doodlaoz maH>Ho>oH odOHuodom wmouo How wounom wdH>Hun was :OHuo>Huu< mo monuodm ovum .ow oHooH 153 mAAanvam comm +m +N m Assess e.uH vo> +mA . +N m nose 0 . so: . H c «H oo> +mA a A N.oA +N mAsmvou - N.m . m.¢H one» +m o.m n o .o +N mAcovoo o.N . N N.NN vs +m m o .m _+m: mAcmVoo H o N o N 0 o $ UQHU +m N.N . c.q swam +m n.m m o m.wH +N mAsovoo o.H .o m.m vo> +m H.o H ¢.hH m.o +N MAao£avoo m m 3a + . I o .mH an A mzv mAhoavoo m . +N N o o o H oAmmzvsm +m . m.m . +N mAconavou N o o w n oAmmzvnm +m as m 0 Oh +N .H m .n +m . e.A . +N +m o o q.o w a onovam . .A + I so A A s N.NA debug w.N o o o.e ox uamuuswom m o.H Hos H H- w.H HOE Hoax NH 0 in e n (6a) ’ " 12 corr 12 corr ’ exp q since AG° W 12 1. 1 AGs "' 8 (67"? (9) 156 Estimates of AC: obtained from eqn (9) for reactions with suitably large values of the quadratic driving force term 0. 5a oAG12 (:>l kcal mol-1) are also given in Tables 18 and 20. For most reactions involving two aquo reactants AG: ~v6-9 kcal mol 1 (Table 18). For reactions involving only one aquo reactant, AC: ~v4-6 kcal mol“1 (Table 20). 2. Entropies and Enthalpies of Activation In view of the marked differences between the experimental free energies of activation and those calculated from the conventional form of the Marcus cross-relation as expressed in eqns (1) or (2), it is of particular interest to compare the behavior of the constit- uent enthalpies and entropies of activation with the predictions of the Marcus model. Expressions similar to eqn (2) can be written for the corres- ponding entropies and enthalpies of activation (134,147): Asz 0.5(AS*1+ AS* 2)(1-40I2 ) + 0. 5(l+2a)ASh (10) 011:2 0.5(AH*1+ AH*2) (1-40:2 ) + 0. S(1+2<:I)AH12 (11) where a is defined by eqn (2a), Asz and Ale are the activation entropy and enthalpy for the cross reaction, A811 and A832 are the activation entropies for the constituent self-exchange reactions, AH* and AH;2 are the activation enthalpies for these self-exchange 11 reactions, and ASi2 and AH;2 are the entrapic and enthalpic driving forces for the cross reaction. Since it is conventional to compute 157 entropies and enthalpies of activation from rate data by using the pre-exponential factor kBT/h rather than 2, the quantities AS* and AH* which appear in eqns (10) and (11) are related to the former quanti- ties, AS* and AH*, by (134,147) AS* 3 AS? - R ln(hZ/kBT) + 0.5R (12) AH* - AH¥ + 0.5RT (l3) Eqns (10) and (11) can be written in a form compatible with eqn (6) by noting that AS* and AH* are reorganizational parameters that may differ from the "Debye-Hfickel corrected" entropies and * * enthalpies of activation, (AS )corr and (AH )corr’ W W * . * W A88 and Ana, respectively, such that (AS )corr A5 + A38 and W' W W * . * . (AH )co AH + AHS. (Thus S8 and H8 are the entropic and by the work terms enthalpic components of S2, the solvent-related work term described in the previous section.) From the preceding expressions, one can write W W 2 * - a * * - - (Aslz)corr ASs o'5[(Asll)corr + (A822)corr 21583“1 4a ) + 0.5(1+2a)ASi2 (14) W W 2 * _ u * * - (AH12)corr AHs 0'5 [(Anll)corr + (AH22)corr ZAHs](1.4a ) + O.5(1+2a)AHi2 (15) W where a is given by eqn (6a) and -AGs - -AH: + TASZ. Since relatively large values of AS: were required in order to fit the experimental 158 free energies of activation with eqn (6), it is of interest to compare the predictions of eqns (14) and (15) with the experimental entropies and enthalpies of activation, respectively, in order to ascertain if the major contribution to AS: arises from.AS: or AHZ. Tables 21 and 22 list the cross reactions considered in Tables 17-20 for which activation parameters have been determined. The listed values of (AH were obtained from the experimental 12)corr quantities AH* by using eqn (13) to yield (AHi'z)app and then by assuming that (AHfz) -(AH (ASW)DH. This presumes that corr 12)app the coulombic work of forming the collision complexes from the sepa- rated reactants is entirely of enthalpic origin at the high ionic strengths encountered. The support for this assertion is that the dependence of the rate constant upon ionic strength for a number of outer-sphere redox reactions has been found to be due almost entirely to the enthalpic term (81,144b). The listed values of (AS*)corr were therefore obtained directly from the conventional experimental quantities AS$ by using eqn (12) (which simply involved the addition of 10 e.u. to AS*). and (AS*1) for the self- Activation parameters (A311)corr corr exchange reactions of interest were given in Table 16. Since some of these quantities have not been directly determined, it was necessary to estimate them from the listed values of (AG11)corr° It was assumed 3+/ 2+ 3+/ 2+ 3+/ 2+ 3+/ 2+ that (Asll)corr- -15 e.u. for Rua aq Euaq Cra md Yba aq since this value has been observed for both FeBZI/2+8 md Viz/2+ self- 4+/3+ 4+/3+ exchange. For Ua aq and Npa aq exchange, (AS was taken as 11)corr 159 Table 21. Thermodynamic Parameters for Some Cross Reactions at 25°C. o a o b o c ”AG12 ‘AH12 AS12 Oxidant Reductant kcal mol- kcal mo1- e.u. 3* Fe2+ 27.2 27 o aq aq 3* c:2+ 53.9 55.5 -5 aq aq 3* + 58.9 60 -3 aq aq rb3+ 2* 11.9 9.8 7 aq aq Fe3* + 22.5 20.7 6 aq aq Fe3+ Eu2+' 25.9 27.4 -5 aq aq Fe3+ 3* 31.7 33.2 -5 aq 89 4+ + Npaq aq 9.1 4.6 15 Eu2+ 3.5 6.8 -11 aq aq + 3+ 3+ + Ru(NH3)6 aq 6.8 12.5 -19 3+ 3+ Ru(NH3)6 Npaq -2.3 7.9 -34 3+ + Ru(NH3)6 aq 16.0 25.0 -30 3+ 3+ Ru(en)aq Npaq 0.5 12.2 -39 3+ 4- Ru(en)3 aq 18.8 29.3 -35 3+ + Ru(NH3)5py aq 12.7 19.0 -21 Ru(NH ) pya+ Euz+ 16 1 25 7 -32 3 5 aq ' ° Table 21. (cont.) 160 o a o b o c ”A612 'AH12 AS12 Oxidant Reductant kcal mol-1 kcal mol-1 e.u. 3+ 2+ Feaq Ru(NH3)6 15.7 8.2 25 Fe3+ Ru(en)2+ 12.9 3.9 30 aq 3 3+ 2+ Feaq Ru(NH3)5py 9.8 1.7 27 3+ 2+ Co(phen) Ru(NH ) 7.5 6.3 4 3 3 6 3+ 2+ Co(phen)3 Ru(NH3)Spy 1.6 -0.2 6 Co(en)3+ Yb2+ 22.2 25.5 -11 3 aq 3+ + Co(phen)3 aq 14.3 18.8 -15 3+ + Co(bpy)3 aq 12.6 17.1 -15 3+ 2+ Ru(bpy)3 Feaq 12.4 25.3 -43 8Free energy driving force for the indicated cross reaction. Values b C taken from Tables 17-20. Entropy driving force for cross reaction, determined from the listed AG° and AS° O - O O 12 12 values by using Ale A612 + TAS 12' Entropy driving force for cross reaction, determined from.the reaction entropies for the appropriate redox couples (Table 15) by . _ 0 red _ o ox . red 0 ox using A812 (Asrc) (Asrc) , where (Asrc) and (Asrc) the reaction entropies for the couples undergoing reduction and are oxidation, respect ively. 161 0N- N.N.. Ha NNI 9H- 36 Hz. cammzvé +m= +N vs 0 m I o o I. o c an mm o N N NH mm o N m 3H +m z +NH :55 . . . . S o N «N N N .c a NN N o N N +N> +NA 56% cm um NN- H.m Ac.HH N I «a .H.HH in: +N> vm cm ON- «.3 NSH oN- o.HH oNH +N=m +N> O O 0 O U“ V” H. m N N NH .HHI o a o S +N> +3.5 MHI N. I . I . . was com H N N NH . N H N 2.. +N in .H «HI 0 . H . NH- o . N . Sam H5... N m N N N +N 2H HHI o.N .m oHI . N. vm> vmwm N .c N N N in I o o I. 0 cm dag HUGO NH 0 N H N OH .H e n N +N .H + o .l o o a.“ V“ H N o o 0H N H N m +m= +N8 O l .0... o + .G . UNHU UQOU N H H N N o N +N +N a I .o .m I o. . H comm coco N N NH N N H +N +N .=.o IHoE Hoax HIHoa Hoax .=.o HIHoaN Hoax HIHoaN Hoax uamuosvom uauvon on? NH uuoo uuoo uuoo UHQUAN« mmHuu< mo comHHomaoo .huooza macaw: aoum voumHaono .NN «Hana 162 | O ' O n' O 'I I %Q NH N N N N NN N o H N +NmN +NH Nvam an m | o 0 II o o . ha HN o N m m MN H m N CH +N> +mA avoo co m | o o | o o A“ «N m o o m mH c m N m +N> +mAamn you - . . - . - . an N Nisms oN m N N m Nm q o N m +Nn +m oo .I. o o .l o o ham m m AH NH N N N N N o N N N +N A szaa +NH=0N sou II o 0 II o 0 Q m m Q NH N H N N NH N N N N +NA szzm +NAam; Voo I O o I o 0 kg“ m .0“ m m q e m 0H e c o n +N A mzvsm +mmm I . . I . . mAcmv woo w c e N m NH H N o o +N 2m +m m 0 o O o m a“ o I m N m ¢ NHI o N m m A mzvnm +mmm vm n m I 0 II o | o II 0 %Q «N H H o 0 AN < o o A +N=m +m A mzvam um n m I . I . I . I . ma HN H o H 0 MN ¢ 0 m o +N> +m A mzvam oNI o.MI N.q HNI m.0I N.m 00: mAcovsm +m +m QMI o.H «.NH oeI m.o m.NH amnz mAcovam +m +m .=.o HIHoaNH Hoax HIHoaN Hoax .=.m HIHoEN Hoax HIHoEN Hoax acouoocom ucmuon Huoo NH uuoo uuoo HUHQUAN¥ Mmcorr’ (AH12)corr’ and (Asz) are listed alongside )calc 2 w-o C01“! , (AHi2)ca1c, and (As: calc. The * the calculated parameters (AG1 w-o 2)w-o calculated values were obtained by inserting the appropriate self- exchange parameters from.Table 16 into eqns (6), (15), and (14), setting As: and AH: equal to zero, and determining a from eqn (6a) by assuming that AG: - O. 165 Inspection of Tables 21 and 22 reveals that when the cross reactions have small values of A812 and A312’ good agreement occurs * between the experimental quantities, (A512)corr and (AHlZ)corr’ and )calca )calc the corresponding calculated parameters (AS* W2 nd (AH"‘1W_2 Naturally the small AH° and A3;2 values for these reactions yield 12 small A612 values, and the good agreement between the calculated and experimental entropies and enthalpies dictates a similar relation between (AG 12)corr and (AG*12)calc. As the magnitude of the free energy driving force increases, the progressively smaller values of calc * (AG12)w-o that are seen relative to (Aclz)corr are typically reflected in values of (A1112)calc that are markedly smaller than calc (Ale)corr and values of (A812)m that are less negative than (A812)corr' It remains to be seen if the difference between experimental and calculated activation parameter can be traced to an enthalpic or 'W entropic work term. The determination of AH8 and AS: is more dif- ficult than that of AS: due to the greater complexity of eqns (14) and (15) relative to eqn (6). However, it is possible to fit the predictions of eqns (14) and (15) with the experimental quantities (Asz) and (AHfz) by trying various values of A32, AHZ, and (301'! 601'! ASS. This procedure will be effective only in those cases where the work terms have a significant effect on the values of the calculated activation parameters. Inspection of eqns (l4) and (15) shows that this will be true when AG° 12 significantly greater than zero (eqn (6a)). Table 23 lists the is sufficiently large so that a is 166 QHI W o H U“ n m NN- N.N N.N NN- N.N- N.N +N=N +NNN A mzvam N m” m- E H..- . .. . z. m N o N N +N: +NHcaN=N N NH- NHN . . . Na N N NN- H H- H N NN- N H- N N +N= +NH mzvam N H I m.m . . as am NN- o o N NH- N H N.N +N: +NmN N OHI ¢.¢ vs vs NH- N.N N.N NH- N.N N.N +NNN +NmN N m I m.q vs vs NH- N.N N.N NH- N.N N.N +N> +NNN N NN- NHHN . . . NN Nmo NN N N o HH NH N H N N +N= +N 0 N NN- m- H .N . . Nmuo HUNo H N N N N N +N +N 0 N MHH Mum N.N NH- N. .HH Nam Nmoo N N +N N +N H .:.o HIHoE Hoax HIHoB Hoax .=.o HIHoa Hoax HIHoa Hoax ucmuosvmm quVon NH uuoo uuoo uuoo wOHMUA ¢mHun swung wcH>om mGOHuomom mmouu maom pom unauoeuuom coHuo>Huo< .mN oHan 167 m m m< n u< Haas .3 3 waHabmmw scum muHsmou moo HQBOH unu «HHga .3m woman one .Amwv can AeHv macs scum voaHmuno .soHuommu muons you :oHuu>Huom mo hmouuco umuMHaonow .Auxwu oomv Mm Imv< ou MOHuaasmmo osu wcchmno was maOHuoavo seam ocu wch: .3 kumHaoHou who? mmon> HosoH on”. mHHHHa .3m<.H.II3u< uqsu mafia—mam .3 Aug cam AmHv 250 60.3 39330 mH moo Home: uSu .voumHH mosHm> mo uHmn sumo you .cOHuomou woouu now GOHum>Huom mo hmHmnucm vmumHsonuo .Auxou oomv msoHuomou oovqusooao: How Hoa Hmux m I moq umnu was msoHuomou oavmloacm now a HI NH 3. uuoo uuoo HI Hoa Hoax o I Mo< uoau m:0Hu gamma on» was HN oHan Bonn oo< .eH uHama scum AN NUHuuo mo Amused swam wouoHsonov .ANN oHnma soumV :OHuo>Huom mo Naouucmo .ANN oHawa aoumv 60Huo>Huoo mo hdesuoo :vouuouuoo HoxuwmIohnon:p .AON can mH moHnoa soumv :OHuw>Huom mo hwuoco swam :wmuoouuoo Hoxummlomnmn:m wHI m o m m MNI a CI H o +Nmm +mA avam MH mHI N.m an m HN- N.N N.N NN- N.N- N.N +NNN +NNamvoN NH .V .- Q..v m. Um ' o 0 II o o 8 m NH N N H N NH N N o N +NH Nam +N N HH m I NIH o m Um MHI m.H o.m NHI o.N m.m A mzvam +mmm OH .:.m HIHoa Hmux H Hos Hoax .=.o HIHoa Hoax HIHoa Hoax acouoavum uammeo NH NH NH uuoo NH uuoo NH uuoo NH NonuH «NNN muHNoHE Ev coHuoN NNN u A may N N NNV A NNN N.N:ouN .NN NHNNN 168 cross reactions from Tables 21 and 22 for which this condition holds. Two sets of activation parameters are given for each system in Table 23. The first set contains experimental values that are taken from Table 22. The other activation parameters in Table 23, labelled (AG*12)calc, (AH*12)calc, md (AS*12)calc, are determined from eqns (6), (6a), (15), and (14) by including trial estimates of the apprOpriate work terms, AGE, ABE, and A82. For the aquo-aquo reactions, AG: was taken to be 6 kcal mol-1, whereas for the nonaquo-aquo reactions AG: was chosen to be 5 kcal mol-1. (The larger AG: value for the aquo—aquo reactions is in keeping with the trends found in Tables 18 and 20.) The estimates of (AG calc which result from the inclusion 12) of these work terms are seen to agree much more closely with (Acf2)corr calc calc than did the (AGf2)F calc values listed in Table 22. Pairs of (AH*12) and (Asz) values are given for each reaction in Table 23. The W upper values were determined by assuming that AG - -TA8:, and the lower values by assuming that AG: - AHZ. These results are also given in graphical form in Figures 5 and 6 which are plots of (AHlZ)corr calc calc versus (AHfz) and of (A812)corr , respectively. Apart from the 003+ - Fe 2+ and Co3+— Cr2+ reactions, it is seen that aq aq aq aq versus (Asz) considerably better agreement between the experimental and calculated activation parameters is obtained using the former assumption (i.e., the closed rather than open points in Figures 5 and 6). Essentially the same results were obtained by choosing different values of AG: within the range of the estimates given in Tables 18 and 20. Reasonable concordance was also obtained by assigning both entrapic 169 Figure 5. Experimental (AH; for cross 2)corr reactions corrected for Debye—Huckel work terms plotted against (AHiz)calc . * calc Key. (AH12) Data are taken from Table 23. The labels for was calculated from eqn (15). each point follow the nmmbering scheme in Table 23. The calculated activation parameters were obtained by assuming that AGZ=m-TASZ (closed circles) and AG: - AH: (Open circles). Reactions 1-3 involving Congz+ were omitted for clarity. The solid straight line has unit slope and passes through the origin. 170 C>O 13 6— 06 T 4 3 E '5 C) x 2. 2 o 9.. *9. I Q 0 -2 l ' l l J l I ‘2 O I 2 (AHE)corr , kcal mol" Figure 5. 4 171 .m ouamHm on how onu :H am>Hw on sun mHHmuon wonuo .AeHv can aouw wouMHoono was Nmm1 (8) 188 where Bred and on are the formal potentials of the species under- f f going reduction and oxidation, respectively. Table 24 summarizes these K12 values. Table 25. Self-Exchange Rate Constants for Selected Noncryptate Couples R11 -1 -l Redox Couple Q! s ) V3-I-/2+ 2 x 10-3 aq E 3+]2+ 1 x -5 uaq 10 Ru(NH3)2+/2+ 3 x 103 The analysis of the Co(NHs)‘:+ - Eu(2.2.l)2+ reaction 13 more complicated. An accurate formal potential for the COh(2.2.2)2+ were made by combining a ten-fold excess of the cryptand with an appropriate salt of the metal ion in 0.1 M_Et4NClO4. The cyclic voltammograms of each of these cryptates were obtained using a HMDE. All of the current-voltage curves had large, sweep-rate-dependent peak separations. For example, the peak separations for Pb(2.2.2)2+ were 530 mV and 490 mV for scan rates of 200 and 50 mV/sec, respec- tively. The smallest peak separations (250 mV at a sweep rate of 200 mV/sec) were found for Tl(2.2.l)+. Even with these large peak separations, it was easily seen that the voltammetric waves of the cryptates were shifted to more negative potentials than those of the free metal ions. These results indicate that the metal ions are stabilized by cryptate formation, although these effects cannot be quantified since thermodynamic information cannot be extracted from the irreversible cyclic voltammograms. 205 The electrochemical irreversibility of these cryptates stems from the involvement of the metallic state in the electrode reactions. Two mechanisms are available to the electroreduction of the cryptates. The so-called EC mechanism consists of the electrochemical reduction of the cryptate (e.g., Pb(2.2.2)2+ + 2e' + Pb(2.2.2)°) followed by a chemical step in which the reduction product dissociates to yield the free cryptand and a metal amalgam. The other possibility is the CE mechanism in which the chemical step precedes the electro- chemical one. Thus, the dissociation of the cryptate gives the free ligand and metal ion, and this last species is then reduced to the amalgam in a subsequent step. An explanation of the observed voltammetric behavior in terms of the EC mechanism could face some difficulties, especially if one makes the assumption that the electroreduction of the cryptate yields a metal atom inside the ligand cavity. Given that the atomic radii of Pb, T1, and Cd are 1.80, 1.90, and 1.55 A, respectively (191), it is unclear how the cavities of the (2.2.1) and (2.2.2) cryptands are able to accommodate these atoms. The application of the EC mechanism becomes more tenable if some modifications are made in the structure which is assumed for the reduction product. It is conceivable that the activation process for the electrode reaction involves a distor- tion of the ligand to a point where the metal ion is no longer com- pletely encapsulated. Such an arrangement would allow the metal ion to be reduced without constraining the resulting metal atom to fit inside the cryptate cavity. The reduction product in this case 206 would be some sort of loose association between the metal atom and .ligand which would fall apart easily in the following chemical step. The ligand reorganization in this mechanism would give rise to a sub- stantial inner-shell component to the activation barrier to electron transfer. The large activation energies for the electrode reactions would then be reflected in slow heterogeneous charge-transfer kinetics and electrochanically irreversible behavior. The slow dissociation rates of the cryptates pose some problems to the application of the CE mechanism to the present electrode reactions. For example, the stability constant of Pb(2.2.2)2+ is 5 X 1012 (15). From Table 8, it is apparent that most divalent (2.2.2) cryptates have formation rate constants of about 104 - lOs'Mfl s‘l. Thus, the dissociation rate constant for Pb(2.2.2)2+ is in the neighborhood of 10.9 s-l. This seems to indi- cate that the dissociation of the cryptate (i.e., the first step of the CE mechanism) occurs too slowly to be seen on the cyclic voltammetric time scale. However, it must be remembered that the estimates made above apply only to bulk solution. The strong electric fields in the interfacial region may bring about field- assisted cryptate dissociation rates which are compatible with the observed electrochemistry. The slow cryptate dissociation rates and the CE mechanism can also be reconciled if one assumes a somewhat different chemical step. Instead of the complete dissociation of the cryptate, this step would now consist of the partial removal of the metal ion from 207 the cryptate cavity. The rate of this step or the reduction rate of the intermediate formed in this step would have to be slow enough to account for the observed electrochemical irreversibility, but at the same.time, these rates would be much faster than the extremely slow dissociation rates of the cryptates. It should be noted that this new CE mechanism is rather similar to the modified EC mechanism dis- cussed earlier. The main difference is that the distortion of the cryptate yields an activated complex in the latter mechanism while it produces a reaction intermediate in the former. At the present either mechanism appears to be adequate for explaining the observed voltammetric behavior. At the outset of this work, there had been some hope that the formation of a cryptate would stabilize an uncommon oxidation state of the encapsulated metal ion. This proved to be the case in the 2+f+ couple. On the other hand, no evidence for the exis- Cu(2.l.l) tence of Cd(2.2.l)+ or Cd(2.2.2)+ was found, although Cd(I) state has been seen in the solid state (192). Evidently the stabilization which Cd(I) would gain through complexation with a cryptand is not sufficient to overcome the strong tendency of this oxidation state to dispro- portionate to Cd(II) and Cd metal. B. The Tl(2.2.2)+/T1(Hg) Couple as a Probe of Cryptate Thermodynamics in anaqueous Solvents 2+ and Pb2+ While the amalgamrforming electroreductions of the Cd cryptates were irreversible, the corresponding reaction for Tl(2.2.2)+ was found to be electrochemically reversible in a variety of solvents 208 (namely water, acetonitrile (MeCN), dimethylsulfoxide (DMSO), N,N-dimethylformamide (DMF), and methanol). Cyclic voltammograms of solutions containing Tl+ and at least a ten-fold excess of (2.2.2) had sweep-rate-independent peak separations of about 60 mV, which is quite close to the value expected for a reversible one-electron process (92). Reversible voltammetric behavior was also obtained for the Tl(I)/T1(Hg) couple in the absence of (2.2.2) in each of the five solvents listed above. In all cases, the voltammetric half-wave potential 355'“33 obtained from the mean of the cathodic and anodic peak potentials. The dependence of Era upon the cryptand concentra- tion was found to be in accordance with the Ligane equation, which applies to the reversible amalgam-forming reduction of labile com- plexes of high stability (193): 3112C - 211']: = -(RT/F) (1n x31 + m 1n [c]t) (1) In eqn (1), E3: and Big/:0 are the half-wave potentials of the Tl(I)/T1(Hg) couple in non-complexing media and in the presence of a large excess of cryptand, respectively, [C]t is the total (analytical) concentration of (2.2.2) cryptand, and'K:1 is the stability constant of the thallium (I) cryptate, TlCm. 10 Data for the dependence of Hie on [C]t were analyzed using eqn (1). It was found that m - l in all solvents; i.e., Tl+ and (2.2.2) form a 1:1 complex, as expected. A value of log K:1 - 6.4 was obtained in aqueous media (u - 0.05) at 25°C. This agrees quite well with an earlier value of 6.3 which was determined using a potentiometric titration (6). 209 The reversible electroreduction of the Tl(2.2.2)+7Tl(Hg) couple indicates that the exchange between Tl+ and Tl(2.2.2)+ is rapid on the time scale of cyclic voltammetry. This finding is in harmony with the rapid exchange rates obtained from NMR.measurements (3,19). It is noteworthy that the previous electrochemical study of T1(2.2.2)+ found irreversible behavior for this complex in propylene carbonate (34). Only a stoichiometric amount of the cryptand was present in the experiments described in reference 34. These condi- tions render the interpretation of the results difficult since the conventional treatment of labile complex-forming systems requires the ligand to be present in a large excess relative to the electro- active metal ion (11). In the present investigation, the concentrae tion of (2.2.2) was always at least ten times that of Tl+. From the form of eqn (1), it can be seen that the Tl+/T1(Hg) and Tl(2.2.2)+/T1(Hg) couples provide a direct means of monitoring the activity (or concentration) of free cryptand. Systems of this kind can be classified as electrodes of the second kind and have been labeled "metal complex" electrodes (194,195). Such couples can be combined with competitive complexation methods to allow the determina- tion of the stability constants of the complexes of electroinactive cations (194,195). A method of this type can be employed to determine a wide range of stability constants by using the following procedure. The half-wave potential, Eyé, for the T1(I)/T1(Hg) couple was deter- mined in solutions containing 0.5‘mMLTl+ (E£;)’ and then after suc- cessive additions of 5-10 m_M_ (2.2.2) cryptand (hf/:0) and 15-50 mg; of 210 the electroinactive metal cation M (Kaine). The ionic strength was maintained at 0.05 or 0.1 M using tetraethylammonium perchlorate (TEAP). Reversible cyclic voltammograms were obtained in each case, at least when M was an alkali metal cation. In general, it was observed that 45?]: < -E¥/:MC < -E1i./:C. The proximity of 3‘12”" to FE]: or E323 was dependent on how effectively M could compete with Tl+ for the cryptand. In order to determine K2, the stability constant for the M(2.2.2) complex, it is necessary to know the equilibrium concentrations of the metal cryptate [MC], the metal ion [M], and the free cryptand in the presence of M, [C]M. As outlined below, [C]M can be determined from the electrochemical data. Knowing [C]M, the remaining terms [MC] and [M] can be found directly since [MC] 8’ [Clt - [C]M and [M] - [Mlt - [MC], where [Mlt is the analytical concentration of M. (Strictly speaking, [MC] - [C]t - [C]M - [TIC], but the last term is usually quite small relative to the others and its neglect represents an error of 3-52. K31, and the analytical If necessary, [TlC] can be calculated from [C]M, concentration of Tl+, [T1+]t') The determination of [C]M can be accomplished by noting that Buck's equation for the effect of complexation upon reversible voltammetric waves (196) can be written for the competition experiment as Tl K [C] +1 TlMC T1 = _ RT Tl _ RT 3 M 1331/2 - E1,2 F ln(I(8 [C]M) F in T1 (2) K3 [C]M (The activity coefficient terms are omitted from eqns (1) and (2) and 211 elsewhere for simplicity.) Although Buck's equation normally only applies to the case where the free ligand is present in a large excess (197), it should apply in the present case even though the free ligand concentration [C]M is typically less than [T1]t. This is because the cryptand concentration will be "buffered" provided that [C]t >’.[Tl]t and the equilibrium M + C ZIMC is maintained during the cyclic voltammogram. (If this is not the case, then the shape of the voltammogram will be severely distorted since the concentration of cryptand increases during the cathodic part of the cycle.) Conse- quently, by obtaining Kgl from eqn (1) in the absence of metal ion M T1 and then inserting K8 into eqn (2), one can determine [C]M and hence K12. In particular, it is interesting to note that when KEITC]M > 1 (i.e., when the stability of the thallium cryptate is sufficiently greater than that for the competing metal cryptate so that Tl TlMC E - E 2 92 95 which allows eqn (1) to be subtracted from eqn (2) to yield 100 mV), the last term in eqn (2) can be neglected, 3],?“ - E12": - 531nm]: — ln[C]M) (3) In this case, [01M and K? can be determined directly from eqn (3). This procedure does not require a knowledge of K21. Therefore, under these conditions, the Tl(I)/T1(Hg) couple responds only to the decrease in free cryptand concentration which results when an excess of metal ion M is added. It is noteworthy that K21 is not needed for a determination of K: under these circumstances. In nonaqueous solvents, it is possible that the small concentration of Tl+ (0.5 mg) is preferentially complexed by water and other impurities, which could 212 lead to false values of Kzl. These incorrect Tl stability constants will not affect the Kgrmeasurements unless the impurities vary in con- centration upon the addition of M, or if they are capable of binding significantly to the much higher (10-50 mg) concentrations of M that were present in these experiments. It is also important to note that [C]M can be calculated using eqn (2) or (3) even when [C]M < [C]t, and that the method is generally capable of allowing K: to be obtained ‘ even when R: < K21. In fact, the practical lower limit of K1: that can be determined using this method is simply that which corresponds to appreciable cryptate formation at the concentrations of M and C chosen. Since we find that K21 lies in the range 106 - 1010 in the solvents studied to date, the method described here allows values of K: in the range ca. 102 to 1010 to be accurately evaluated. (It is difficult to obtain K: values when K: >‘K:1 since the T1+ ceases to be an effective competitor. In practice, the difference between T1 TlMC Eyh and E55 becomes too small to be measured reliably.) Table 27 lists some preliminary results for K: in DMSO and MeCN. It is seen that good agreement is obtained between the present and earlier determinations of K? for Cs(2.2.2)+ in DMSO (198). Potassium and thallium cations were uniformly found to exhibit the largest values of K: in each solvent, which is in harmony with the judicious fit of these bare cations into the (2.2.2) cryptate cavity. This simple electrochemical method therefore appears to provide a powerful approach to the exploration of the thermodynamics of cryptate and other macrocyclic complexation in a range of solvents. 213 Table 27. Formal Stability Constants (log Ks) for Univalent (2.2.2) Cryptates in Dimethylsulfoxide ' and Acetonitrile at 25°C. Cryptate DMSOa MeCNb Li(2.2.2)+ < 1 7.3 Na(2.2.2)+ 5.3 10.6 1<(2.2.2)+ 6.0 11.1 Cs(2.2.2)+ 1.6 -- Tl(2.2.2)+ 6.2 12.3 aDetermined at u - 0.05. bDetermined at u = 0.1. While its requirement of electrochemically reversible systems may seem restrictive, the T1 probe does offer an advantage over the principal alternative, potentiometry. Because the T1 technique is not a zero- current method, it is less likely to be affected by impurity currents. Thus, the T1 probe can be considered an important complement to the well-established potentiometric methods. CHAPTER VIII CONCLUSIONS AND SUGGESTIONS FOR FURTHER WORK 214 A. Conclusions 1. Transition Metal Cryptates A recurrent theme in the studies of transition metal cryptates is the important influence of ion hydration. The relatively low stability constants of the trivalent Eu cryptates result from the enthalpic destabilization associated with the extensive (or at least partial) desolvation of the aquo cations upon their encapsulation within the cryptate cavity. The surprising kinetic stabilities of these trivalent cryptates stem from the large changes in solvent structure which are needed to attain the transition states in the dissociation reactions. One of the primary reasons for the enhanced 3+] 2+ is aq a smaller solvent-reorganization barrier to electron transfer in the redox reactivities of the Eu cryptate couples relative to Eu former species. The comparatively weak hydration of the trivalent cryptates could well be the explanation for the tendencies of these ions to associate strongly with F- and OH-. The variations seen in the standard potentials of the Eu(2 .2 .1)3+/2+ 3442+ 3+/ 2+ , Eu(ZB.2.l) , and Eu(2.2.2) couples indicate that the structure of ligand can have a strong influence on the rela- tive stabilities of the encapsulated Eu(III) and Eu(II) oxidation states. Therefore, it would seem that the apprOpriate alterations in ligand structure could yield a cryptate couple which has a specified oxidizing or reducing ability. The cryptands could be modified by replacing some of the ether oxygens with nitrogen or sulfur donor 215 216 atoms, by binding various functional groups to the ethylene bridges, by lengthening or shortening these bridges, or by introducing substi- tuents into the aromatic ring of (23.2.1). This last approach could yield a series of structurally similar redox couples with varying standard potentials. These cryptates could be useful in studies of homogeneous electron-transfer kinetics since the intrinsic components of the activation barriers of these complexes could well be reasonably constant while the driving-force components obviously have to vary. One could then investigate the driving-force dependence of homogeneous cross—reaction rate constants. A.study of this kind could also be undertaken with structurally dissimilar reactants, but some sort of model would have to be employed in order to account for the effects of the disparate intrinsic terms. The rate data from Chapter VI appear to support the analysis of Chapter V which suggested that electron transfer involving f orbitals was not especially nonadiabatic. The cross-reaction rate constants in Chapter VI conform to the Marcus cross-relation within the limits of experimental uncertainty. Thus the electron-transfer reactions of the Eu cryptates do not show any indications of nonadiabaticity. In 3+] 2+ fact, if the Eu 8Q couple were nonadiabatic, one would expect the cryptates to be even more so since the encapsulation of Eu(III) or Eu(II) inside the cavity of a large organic ligand should signifi- cantly diminish the already low probability for electron transfer within the activated complex. (In other words, one would anticipate that the transferred electron would have to tunnel over a greater distance in the cryptate reactions.) On the basis of the present evidence, the Eu(III)/(II) couples in this study are no less 217 adiabatic than couples which transfer electrons to and from d orbitals. 2. Reaction Entropies The effects of hydration also pervade the reaction entropy study. The large As;c values for M(III)/(II) aquo couples are due to the extensive solvent structure which forms around the trivalent ion as the result of field-assisted hydrogen bonding involving the protons on the aquo ligands. The lower As;c values of the ammine couples presumably result from.the weaker hydrogen bonding between the ammine . protons and the surrounding solvent. Thus, interactions between the ligands and the solvent play a major role in determining the Asgc for a transition metal couple. Under certain circumstances, the ligand effects predominate to such an extent that the reaction entrOpy of a couple can be obtained simply by considering individual contri- butions from each ligand in the complex. For example, the similar As;c values of the Fe(bpy)3+/2+ )g+/2+ 3)g+/2+ and Os(NH3):+/2+ couples can be explained in these terms. and Ru(bpy couples or of the Ru(NH Perhaps the most graphic instance of the ligand influence is in the 3+/ 2+ 3+/ 2+ 3+/ 2+ 3+/2+ series, where the As;c values can be obtained by adding together . Ru(NH3)2(bpy) . and Ru(bpy) separate components for each bpy and NH ligand. However, the 3 reaction entropy is not always solely dependent on the ligands in the redox couple. The different Asgc values for analogous Co and Ru couples (see Table 11) indicate that the central metal ion can also have a significant effect on the reaction entrOpy. It would appear that two redox couples will have comparable reaction entrOpies when 218 they have the same ligand composition and when the electronic struc- tures of the two central metal ions are similar. These conditions are restrictive but some useful generalizations might be made. The Ru(III)/(II) systems could serve as models for analogous couples in which an electron is transferred to or from a nonbonding orbital, and the appropriate Co(III)/(II) couples may be representative of those systems where an antibonding orbital is involved in the electron transfer. Unfortunately, only a limited number of metal ion couples are suitable for studying the above assertions, so these will remain largely untested at the present. The reaction entropies provide a convenient means of calculating the entropy of reaction for homogeneous redox processes. These quanti- ties can also be obtained from the temperature dependence of the appropriate equilibrium cell potentials. The potentiometric method has a number of drawbacks. The cell potential of each reaction has to be measured separately. In contrast, a relatively small number of As;c values can be used to calculate the thermodynamic driving forces of a host of homogeneous reactions. The potentiometric method frequently requires the maintenance of stable concentrations of reactants and products. This can be difficult considering that many of the species in these reactions are unstable with respect to oxi- dation or aquation, especially at high temperatures. In the As;c method, these species only need to be present for the duration of a cyclic voltammogram.(i.e., a few seconds). Finally, the uncertain- ties in the os;c values which stem from extrathermodynamic assump- tions will cancel when the difference between two reaction entrOpies 219 is taken to calculate the ASE2 value for a cross reaction. In essence, these advantages of the As;c method enabled the study per- formed in Chapter V to scrutinize a wide range of reactions. It would have been impossible, or at best laborious, to obtain the driving-force parameters by means of direct potentiometry. The correlation between as;c and AGfl, the activation free energy for self-exchange, proved to be effective in explaining certain patterns in redox reactivity. However, this correlation cannot be applied indiscriminately. It appears to retain general validity when it is employed with relatively simple redox couples, but the problems which are encountered when the correlation is applied to the Eu cryptates may indicate that the correspondence between as;c and 06:1 is not as firm when more complex ligand structures are under considera- tion. Large changes in ligand conformation can occur in the electron- transfer activation process for such systems. These changes will be reflected in As;c only if the alteration in ligand conformation has a significant effect on the ability of the solvent to approach the redox center. Therefore, it may well be that the correlation is only useful to a limited degree for complex ligand systems such as those seen in many biological redox couples. 0n the whole, the us;c - AGE1 correla- tion is best used for interpreting the trends shown by a large number of redox couples (Figure 4 is one example of this use.) The correla- tion is not exact, so one cannot expect consistent success when applying it to pairs or trios of redox couples unless these systems are very similar to each other. B. Suggestions for Further Work Because the hydration of metal cations has emerged as a crucial factor in the cryptate and reaction entrOpy studies, it will be of interest to determine if similar investigations in nonaqueous solvents yield the same sort of results. The different dielectric constants and specific solvating abilities of these solvents should produce a wide range of As;c values. These reaction entropies should contain useful information on the solvent structure and the ligand- solvent interactions in nonaqueous solution. One can also expect some changes in the thermodynamics and kinetics of the Eu and Yb cryptates when these complexes are studied in a nonaqueous solvent. Some increases in the cryptate dissociation rates might be found in weakly solvating media. The difference in the formal potentials of the Eu cryptate and solvated Eu couples will be related to the extent of solvation of Eu(III) relative to that of Eu(II). A number of further investigations on the homogeneous redox kinetics of cryptates could be performed. Information on the reac- 3+/ 2+ tivity of Yb(2.2.l) would be useful for comparison with the Eu cryptates and also for its own sake. Because the Yb(2.2.l)3+/2+ couple has a very negative standard potential, the reactions of this species with most commonly employed redox reagents will have large driving forces. The relevant reactions in Chapter V all had sub- stantial driving forces and at least one aquo reactant. It would be interesting to determine if the reactions between the Yb cryptate 220 221 and some appropriate nonaquo coreactants contain the same solvent- related work terms (06:) found in Chapter V. The presence of the AC: term was ascribed to the solvent reorientation needed to form the highly charged precursor complex. It seems likely that much of the solvent structure in these complexes is induced by the aquo ligands on one or both of the reactants. Because nonaquo complexes are less strongly hydrated, it might be expected that a redox reaction between two such species would have a lower AC: value. Unfortunately, 3+l2+ will cause the extremely negative standard potential of Yb(2.2.l) some difficulties in the determination of the self-exchange rate constants for this couple. The Marcus cross-relation begins to break down when the driving force of a cross reaction is large. This will be the case for the reactions between the Yb cryptate and most well- characterized coreactants. Therefore, one must employ redox couples which also have particularly negative E° values; e.g., U22”+ or Yb3+/ 2+. aq The homogeneous redox kinetics of the Cu(2.l.l)2+/+ couple would be worthy of examination. The low As;c value of this couple and the more rigid structure of the (2.1.1) ligand relative to (2.2.1) and 2”” should (2.2.2) suggest that the self-exchange rate for Cu(2.l.l) be rapid. Studies of the dissociation kinetics of the divalent and monovalent Cu cryptates and the characterization of any anion effects (such as those observed for the Eu(III) cryptates) would be useful preliminaries for the investigation of the electron-transfer reactions of Cu(2.l.l)2+/+. 222 The range of formal potentials shown by the Eu(2.2.l)3+/2+, 3+]2+ 3+/ 2+ Eu(ZB.2.l) , and Eu(2.2.2) couples demonstrates that the Ef value of the Eu(III)/(II) couple can be altered considerably by changes in the cryptand structure. Further insight into this kind of ligand effect could be obtained by determining the formal poten- tials for complexes formed between Eu(III)/(II) and the various polyaza and polythia analogs of the (2.2.1), (2.2.1), and (2.1.1) cryptands. The replacement of the oxygen atoms in the cryptands by nitrogen and sulfur should produce large changes in the relative stabilities of the Eu(III) and Eu(II) complexes. The nitrogen and sulfur donor atoms could well stabilize the Eu(III) state via covalent bonding and modify the ligation and electron-transfer kinetics of the trivalent and divalent complexes. The technique of determining stability constants for (2.2.2) com- plexes by means of a competition with T1+ should prove to be applicable to a large number of solvents. As noted in Chapter VII, this method can be employed when the Tl+/T1(Hg) and Tl(2.2.2)+/T1(Hg) couple are electrochemically reversible and when T1(2.2.2)+ and the other (2.2.2) complexes under study are substitutionally labile. Since T1(2.2.2)+ is quite stable, the Tl+ competition method can be utilized to determine stability constants which are too large to be obtained from spectrophotometric techniques. This aspect of the Tl+ method takes on special significance in those nonaqueous solvents where pH titrations cannot be used to determine cryptate stability constants. Thus, the Ks values for a number of (2.2.2) complexes can be obtained in a range of solvents by means of this T1+ competition technique. 223 The enthalpies and entropies of complexation can then be calculated from the temperature dependences of the stability constants. These studies will furnish useful insights into the influences of the solvent on cryptate stability. The Tl+ method should be easily applicable to univalent cryptates, but some obstacles could be encountered with divalent complexes. As noted in Chapters I and III, a divalent cryptate has a larger stability constant and slower dis- sociation rate constant than the corresponding univalent complex. Therefore, the T1+ ion may not be capable of competing effectively with alkaline earth ions for the (2.2.2) ligand. In other cases, thermodynamic information will not be available since the voltammetric waves will be distorted by the slow ligand exchange kinetics of the divalent cryptate. Despite these possible limitations, the Tl+ competition method represents a simple and convenient means for probing cryptate stabilities in a variety of solvents. LIST OF REFERENCES 224 LIST OF REFERENCES B. Dietrich, J. M. Lehn, and J. P. Sauvage, Tetrahedron Lett., 2885 (1969). B. Dietrich, J. M. Lehn, and J. P. Sauvage, Tetrahedron Lett., 2889 (1969). - ‘ J. M. Lehn, Struct. Bonding (Berlin), ié, 1 (1973). Of particular interest to the present work are the structures of: (a) La(2.2.2)3+ from F. A. Hart, M. B. Hursthouse, K. M. Abdul Malik, and S. Moorhouse, J. C. 8. Chem. Comm., 549 (1978). (b) Eu(2.2.2)3+ from M. Ciampolini, P. Dapporto, and N. Nardi, (c) Co(2.l.l)2+ from F. Mathieu and R. Weiss, J. C. S. Chem. Comm., 816 (1973). (d) Tl(2.2.2)+ from D. Moras and R. Weiss, Acta. Cryst., B29, 1059 (1973). (e) Pb(2.2.2)2+ from B. Matz and R. Weiss, Inorg. Chem., 12) 2094 (1974). A good compilation of cryptate crystal structures can be found in A. I. POpOV and J. M. Lehn, "Physicochemical Studies of Crown and Cryptate Complexes" in "Coordination Chemistry of Macrocyclic Compounds", G. A. Melson, editor, Plenum, New York, N. Y. (1979). J. M. Lehn and J. P. Sauvage, J. Amer. Chem. Soc., 21, 6700(1975). C. J. Pedersen, J. Amer. Chem. Soc.,‘BE, 2495 (1967). C. J. Pedersen, J. Amer. Chem. Soc., 82, 7017 (1967). C. J. Pedersen and H. K. Frensdorff, Angew. Chem. Int. Ed. Engl., ‘11, 16 (1972). 225 10. 11. 12. 13. 14. 15. 16. 17. 18. 19. 20. 21. 22. 23. 24. 25. 226 J. J. Christensen, D. J. Eatough, and R. M. Izatt, Chem. Rev.,Z£_ 351 (1974). An extensive review of crown and cryptate chemistry although somewhat out of date. This and (unless otherwise noted) the other ionic radii in this work are taken from R. D. Shannon and C. T. Prewitt, Acta Cryst., .225, 925 (1969). The values quoted are for a coordination number of six. E. Kauffmann, J. M. Lehn, and J. P. Sauvage, Helv. Chim. Acta,.2, 1099 (1976). J. M. Lehn, Acc. Chem. Res.,‘;l, 49 (1978). F. Armand-Neu, B. Spiess, and M. J. SchwingAWeill, Helv. Chim. Acta,.§Q, 2633 (1977). J. M. Lehn and F. Montavon, Helv. Chim. Acta, 61, 67 (1978). B. Spiess, F. Arnaud-Neu, M. J. Schwing-Weill, Helv. Chim. Acta, 32, 1531 (1979) . G. Anderegg, Helv. Chim. Acta, 58, 1218 (1975). "Handbook of Chemistry and Physics", 56th Edition, CRC Press, Cleveland, 1975, p. F210. J. M. Lehn, J. P. Sauvage, and B. Dietrich, J. Am. Chem. Soc.,22, 2916 (1970). J. M. Ceraso and J. L. Dye, J. Am. Chem. Soc., 25, 4432 (1973). Y. M. Cahen, J. L. Dye, and A. I. Popov, J. Phys. Chem., 12, 1292 (1975). J. M. Ceraso, P. B. Smith, J. S. Landers, and J. L. Dye, J. Phys. Chem., 81, 760 (1977). E. Mei, A. I. Popov, and J. L. Dye, J. Phys. Chem.,.§1, 1677 (1977). K. Henco, B. Tflmmler. and G. Maass, Angew. Chem. Int. Ed. Engl., 16, 538 (1977). V. M. Loyola, R. G. Wilkins, and R. Pizer, J. Am. Chem. Soc.,.gz 7382 (1975). 26.V. M. Loyola, R. Pizer, and R. G. Wilkins, J. Am. Chem. Soc., 99, 7185 (1977). Note that the values for the (2.2.1) and (2.1.I7- systems given in this paper supplant the earlier ones reported in reference 25. 27. 28. 29. 30. 31. 32. 3'3. 34. 35. 36. 37. 38. 39. 40. 41. 42. 43. 44. 455. 46» 227 B. C. Cox and H. Schneider, J. Am. Chem. Soc.,_22, 2809 (1977). B. C. Cox, H. Schneider, and J. Stroka, J. Am. Chem. Soc., 100, 4746 (1978). F. Peter and M. Cross, J. Electroanal. Chem., 53, 307 (1974). J. P. Gisselbrecht, F. Peter, and M. Cross, J. Electroanal. Chem, 2, 447 (1974). F. Peter and M. Cross, J. Electroanal. Chem.,_§l, 245 (1975). J. P. Gisselbrecht, F. Peter, and M. Gross, J. Electroanal. Chem., 13, 315 (1976). A. Louati and M. Cross, Electrochim. Acta,_21, 7 (1976). J. P. Gisselbrecht and M. Gross, J. Electroanal. Chem., 15, 637 (1977). F. Peter, J. P. Gisselbrecht, and M. Gross, J. Electroanal. Chem., 86, 115 (1978). J. P. Gisselbrecht, F. Peter, and M. Gross, J. Electroanal. Chem., 26, 81 (1979). D. Britz and D. Knittel, Electrochim. Acts, 29, 891 (1975). F. A. Cotton and G. Wilkinson, "Advanced Inorganic Chemistry", Third Edition, Interscience, New York, 1972, pp. 208, 1075. G. N. Lewis and M. Randall, "Thermodynamics", Second Edition, McGrawbHill, New York, 1961, p. 399. C. M. Criss and M. Saloman in "Physical Chemistry of Organic Solvent Systems", A. K. Covington and T. Dickinson, editors, Plenum Press, New York, 1973, Chapter 2. R. E. Powell and W. M. Latimer, J. Chem. Phys“ 12, 1139 (1951). J. W. Cobble, J. Chem. Phys.,_21, 1446 (1953). P. George, G. I. H. Hanania, and D. H. Irvine, J. Chem. Phys., 22; 1616 (1954). P. George, G. I. H. Hanania, and D. H. Irvine, Recl. Trav. Chim. Pays-Bas, 15, 759 (1956). K. J. Laidler, Can. J. Chem., 34, 1107 (1956). R. M. Noyes, J. Am. Chem. Soc.,.gfl, 513 (1962). 47. 48. 49. 50. 51. 52. 53. 54. 55. 56. 57. 58. 59. 60. 61. 62. 63. 64. 65. 66. 228 H. S. Frank.and M. W. Evans, J. Chem. Phys., 13, 507 (1945). H. 8. Frank and W.-Y. Wen, Discuss. Faraday Soc., 24, 133 (1957). R. E. Powell, J. Phys. Chem., 58, 528 (1954). M. J. Weaver, J. Phys. Chem., fl, 2645 (1976). J. Lin and W. G. Breck, Can. J. Chem., 43, 766 (1965). H. S. Lim, D. J. Barclay, and F. B. Anson, Inorg. Chem., 11, 1460 (1972). - G. N. Coleman, J. W. Gesler, F. A. Shirley, and J. R. Kuempel, Inorg. Chem., 12, 1036 (1973). C. M. Elson, I. J. Itzkovich, J..McKenney, and J. A. Page, Can. J. Chem., 23, 2922 (1975). T. Matsubara and P. C. Ford, Inorg. Chem., 12, 1107 (1976). J. A. Marchant, T. Matsubara, and P. C. Ford, Inorg. Chem., lg, 2160 (1977). J. Gulens and J. A. Page, J. Electroanal. Chem., 55, 239 (1974). J. Gulens and J. A. Page, J. Electroanal. Chem., 61, 215 (1976). R. G. Wilkins, "The Study of Kinetics and Mechanism of Reactions of Transition Metal Complexes", Allyn and Bacon, Boston, 1974, Chapter 5. L. E. Bennett, Frog. Inorg. Chem., 18, 1 (1973). H. Taube, "Electron Transfer Reactions of Complex Ions in Solution", Academic Press, New York, 1970. H. Taube, H. Myers, and R. L. Rich, J. Am. Chem. Soc., 12, 4118 (1953). H. Taube and H. Myers, J. Am. Chem. Soc., 16, 2103 (1954). For summaries of these studies see: (a) T. J. Williams and C. S. Garner, Inorg. Chem., 2, 2058 (1970), (b) D. W. Carlyle and J. H. Espenson, J. Am. Chem. Soc., 21, 599 (1969). J. P. Candlin, J. Halpern, and S. Nakamura, J. Am. Chem. Soc.,gé, 2517 (1963). J. F. Endicott and H. Taube, Inorg. Chem., 4, 437 (1965). 67. 68. 69. 70. 71. 72. 73. 74. 75. 76. 77. 78. 79. 80. 81. 82. 83. 84. 85. 86. 229 D. Seewald, N. Sutin, and K. O. Watkins, J. Amer. Chem. Soc., 21, 7307 (1969). W. G. Mbvius and R. G. Linck, J. Am. Chem. Soc.,_22, 2677 (1970). N. Sutin, Acc. Chem._Res., l, 225 (1968). H. J. Price and H. Taube, Inorg. Chem., 2, 1 (1968). M. P. Liteplo and J. F. Endicott, Inorg. Chem., 19, 1420 (1971). T. L. Toppen and R. G. Linck, Inorg. Chem., 19, 2635 (1971). R. K. Murmann, H. Taube, and F. A. Posey, J. Am. Chem. Soc., 12, 262 (1957). , W. L. Reynolds and R. W. Lumry, "Mechanisms of Electron Transfer", Ronald Press, New York, 1966. N. S. Hush, Trans. Faraday Soc.,-51, 557 (1961). V. G. Levich and R. R. Dogenadze, Proc. Acad. Sci., USSR, Phys. Chem. Sect.,.133. S91 (1960); Collect. Czech. Chem. Commun.,_2§, 193 (1961). R. A. Marcus, J. Phys. Chan.,_6_7_, 853 (1963). R. A. Marcus, Ann. Rev. Phys. Chem., 15, 155 (1964). and references cited therein. R. A. Marcus, J. Chem. Phys., 43, 679 (1965). H. Taube, Adv. Chem. Ser., 162, 127 (1977). G. M. Brown and N. Sutin, J. Am. Chem. Soc., 101, 883 (1979). R. A. Robinson and R. H. Stokes, "Electrolyte Solutions", Second Edition, Butterworths, London, 1959, p. 230. R. G. Linck in MI? International Review of Science, Inorganic Chemistry, Series 1, Vol. 9, M. L. Tobe, editor, Park Press, Baltimore, 1972. R. J. Taylor and A. A. Humffrey, J. Electroanal. Chem., 42, 347 (1973). D. J. Pruett, Michigan State University, Ph.D. Thesis (1978). B. A. Moyer and T. J. Meyer, J. Am. Chem. Soc., 100, 3601 (1978). 87. 88. 89. 90. 91. 92. 93. 94. 95. 96. 97. 98. 99. 230 E. L. Yee and M. J. Weaver, Inorg. Chem., 22, 1077 (1980). (a) I. I. Creaser, J. MacB. Harrowfield, A. J. Herlt, A. M. Sargeson, J. Springborg, R. J. Geue, and M. R. Snow, J. Am. Chem. Soc., 22, 3181 (1977); (b) A. M. Sargeson, Chem. Br., 23 (1979). F. Bottomley and S. B. Tong, Inorg. Chem., 22, 243 (1974). J. M. Lehn and M. E. Stubbs, J. Am. Chem. Soc., 22, 4011 (1974). M. J. Weaver and F. C. Anson, J. Electroanal. Chem., _6_§_, 711 (1975). See, for example, (a) R. N. Adams, "Electrochemistry at Solid Electrodes", Marcel Dekker, New York, 1969, pp. 145-149; (b) D. T. Sawyer and J. L. Roberts, Jr., "Experimental Electrochemistry for Chemists", Wiley, New York, 1974, Chapter 7. R. S. Nicholson, Anal. Chem., 21, 1351 (1965). M. J. Weaver, J. Phys. Chem., g4, 568 (1980). Strictly speaking, this procedure yields the so-called "half-wave" potential E!” which will equal Ef only when the diffusion coefficients of the oxidized and reduced forms of the redox couple are identical. However, Ehh normally differs from Ef by only 2-3 mV. See Chapter IV for a more detailed discussion. The (23.2.1) cryptand is similar to (2.2.1) except that one of the central dioxyethylene groups is replaced by the analogous catechol (3). Although the determination of A8;c requires at least one extrathermodynamic assumption to be made (see Chapter IV), this uncertainty will cancel when differences in AS° are rc considered as in eqn (2). It is important to note that the time required to attain true equilibrium will be determined by the ligand exchange rate (i.e., the first-order rate constant for dissociation of the complex) irrespective of whether the complex is allowed to aquate spontaneously or the free metal ion is allowed to complex with added ligand. See, for example, D. R. Crow, "Polarography of Metal Cbmplexes", Academic Press, New York, 1969, pp. 73-77. 231 100. Equation (4) is a form of the well-known Lingane equation; it should be carefully distinguished from equation (1) since the former relation will apply when only the oxidized species is complexed to form substitutionally labile complexes. 101. See, for example, K. J. Laidler, "Chemical Kinetics", Second Edition, McCrawbHill, New York, 1965, pp. 88-90. 102. (a) L. Pauling, "The Nature of the Chemical Bond", Cornell University Press, Ithaca, Third Edition, 1960, Chapter 13; (b) S. W. Benson and C. S. Copeland, J. Phys. Chem., 21, 1194 (1963). 103. See, for example, J. P. Hunt, "Metal Ions in Aqueous Solution", Benjamin, New York, 1963, Chapter 2. 104. J. C. Barnes and P. A. Bristow, Inorg. Nucl. Chem. Lett., 5, 565 (1969). 105. M. Diebler, M. Eigen, G. Ilgenfritz, G. Mesh, and R. Winkler, Pure App. Chem., 22, 93 (1969). 106. The estimate, St *‘-10 to -15 e.u., arises from the conven- tional use for bimolecular reactions of the transition-state formalism with the pre-exponential factor ( T/h) exp (ASI/R) rather than the simple collisional model wit the bimolecular collision frequency Z. See, for example, reference 60 or reference 101. 107. F. H. Spedding, J. A. Rard, and A. Habenschuss, J. Phys. Chem., g;, 1069 (1977). 108. For further details, see Chapter 4 and reference 48. 109. For a general review, see J. N. Agar, Adv. Electrochem. Electrochem. Eng.,.2, 31 (1963). 110. For example, see (a) E. D. Eastman, J. Am. Chem. Soc., 50, 292 (1928); (b) A. J. de Bethune, T. S. Licht, and "' N. Swendeman, J. Electrochem. Soc., £22, 616 (1959); (c) A. J. de Bethune, J. Electrochem. Soc., 197, 829 (1960); (d) G. Milazzo and M. Sotto, Z. Phys. Chem. (Ffankfurt am Main),_§2, 293 (1967); (e) G. Milazzo, M. Sotto, and C. Devillez, Z. Phys. Chem., 22, 1 (1967). 111. The values for Hg and Pt are obtained by combining the relative Thomsen coefficients (vs. Pb) quoted in G. P. Harnwell, "Principles of Electricity and Magnetism", MCGrawéHill, New York, 1939; p. 187, with the absolute Thomsen coefficient for Pb given in reference 109. 112. 113. 114. 115. 116. 117. 118. 119. 120. 121. 122. 123. 124. 125. 126. 127. 128. 129. 232 R. S. Nicholson and I. Shain, Anal. Chem., _2, 706 (1964). If the concentration of the electroactive species is known, the diffusion coefficient can be calculated from equation (5) or (6) from Chapter III. C. I. H. Hanania, D. H. Irvine, W. A. Eaton, and P. George, J. Phys. Chem., 1;, 2022 (1967). Reference 39, p. 137. Reference 92b, pp. 29-30. Reference 82, pp. 42, 465. R. R. Buckley and E. E. Mercer, J. Phys. Chem.,‘zg, 3103 (1966). (a) D. K. Lavallee, C. Lavallee, J. C. Sullivan, and E. Deutsch, Inorg. Chem., 22, 570 (1973); (b) C. Lavallee and D. K. Lavallee, Inorg. Chem., 19: 2601 (1977). J. J. Kim.and P. A. Rock, Inorg. Chem., 2, 563 (1969). P. George, G. I. H. Hanania, and D. H. Irvine, J. Chem. Soc., 2548 (1959). See, for example, F. Basolo and R. G. Pearson, "Mechanisms of Inorganic Reactions", Second Edition, Wiley, New York, 1967, pp. 62-64. (a) E. R. Nightingale, J. Phys. Chem., 53, 1381 (1959); (b) T. Takahashi and T. Kolso, Bull. Chem. Soc. Japan, 49, 2734 (1976) . ‘— M. J. Weaver and T. L. Satterberg, J. Phys. Chem., 2;, 1772 (1977). Reference 38, p. 592. H. J. Peresie and J. A. Stanko, Chem. Commun., 1674 (1970). H. C. Stynes and J. A. Ibers, Inorg. Chem., 19: 2304 (1971). N. Sailasuta, F. C. Anson, and H. B. Gray, J. Am. Chem. Soc., 101, 455 (1979). Itrshould be noted that the reaction entropies tabulated in this paper are relative values based on the assumption that As;c - 0 for the H+IH2 couple. N. J. Hair and J. K. Beattie, Inorg. Chem., 12, 245 (1977). 130. 131. 132. 133. 134. 135. 136. 137. 138. 139. 140. 141. 142. 143. 233 N. Sutin, in "Tunneling in Biological Systems", B. Chance, D. C. Devault, H. Frauenfelder, R. A. Marcus, J. R. Schreiffer, and N. Sutin, editors, Academic Press, New York, 1979, p. 201. E. Buhks, M. Bixon, J. Jortner, and G. Navon, Inorg. Chem., lg, 2014 (1979). W. BBttcher, G. M. Brown, and N. Sutin, Inorg. Chem., l§, 1447 (1979). Estimates by means of the Marcus cross-relation from the 3+/2+ rates of outer-sphere cross reactions which involve Cr(OHZ)6 and which have small driving forces (134). M. Chou, C. Creutz, and N. Sutin, J. Am. Chem. Soc., 22, 5615 (1977). K. V. Krishnamurty and A. C. Wahl, J. Am. Chem. Soc.,}§9, 5921 (1958). J. Silverman and R. W. Dodson, J. Phys. Chem.,,éés 345 (1952)- R. C. Young, F. R. Keene, and T. J. Meyer, J. Am. Chem. Soc.,fi, 2468 (1977). F. P. Dwyer and A. M. Sargeson, J. Phys. Chem., 22, 1892 (1961). H. M. Neumann, quoted in R. Farina and R. G. Wilkins, Inorg. Chem., 1, 514 (1968). For recent reviews which deal with this point, see (a) R. G. Linck in "Homogeneous Catalysis", G. N. Schrauzer, editor, Marcel Dekker, New York, 1971, Chapter 7; (b) N. Sutin in "Inorganic Biochemistry", G. L. Eichhorn, editor, Elsevier, New York, 1973, Chapter 19; (c) R. G. Linck in "MTP Inter- national Review of Science", Inorg. Chem. Series II, Vbl. 9, Butterworths, London, 1974, Chapter 7; (d) R. G. Linck in "Survey of Progress in Chemistry", A. F. Scott, editor, Vol. 7, Academic Press, New York, 1976, p. 89. (a) D. P. Rillema, J. F. Endicott, and R. C. Patel, J. Am. Chem. Soc., 22, 394 (1972); (b) D. P. Rillema and J. F. Endicott, J. Am. Chem. Soc., 23, 8711 (1972). I. Bodek and G. Davies, Coord. Chem. Rev., 22, 269 (1974). B. Falcinella, P. D. Felgate, and G. S. Lawrence, J. Chem. Soc. Dalton, 1 (1975). 144. 145. 146. 147. 148. 149. 150. 151. 152. 153. 154 O 155. 156. 157. 158. 159. 234 (a) A. Ekstrom, A. B. MCLaren, and L. E. Smythe, Inorg. Chem., 22, 2899 (1975); (b) A. Ekstrom, A. B. McLaren, and L. E. Smythe, Inorg. Chem., 22, 2853 (1976); (c) A. Ekstrom, A. B. McLaren, and L. E. Smythe, Inorg. Chem., 22, 1032 (1977). G. M. Brown, H. J. Krentzien, M. Abe, and H. Taube, Inorg. Chem., 22, 3374 (1979). Reference 74, pp. 123, 129. R. A. Marcus and N. Sutin, Inorg. Chem., 22, 213 (1975). D. H. Huchital, N. Sutin, and B. Warnquist, Inorg. Chem.,_§, 838 (1967). J. c. Hindman and E. s. Kritchevsky, J. Am. Chem. Soc., _7_2_, 953 (1950). D. Cohen and J. C. Hindman, J. Am. Chem. Soc., 13, 4679 (1952). The isothermal cell results in this reference were converted to the absolute scale by using dsgc - 21 e.u. for the H+IHZ couple. H. S. Habib and J. P. Hunt, J. Am. Chem. Soc., 22, 1668 (1966). The quoted 2 value _isll2 obtained from the expression (79, 81) Z - r2N(8w k3 Tmr’I h°103where r is the distance between the metal centers in) the collision complex and m1. is the effective reduced mass of this complex. Typical values for the present reactants (r- 7 X 10 8 cm and mi. - 100) yie1d2~6><10 oM '- Values of 3 used to calculate (AGW) from eqn (5) were obtained from the sum of the radii o the reacting ions. M. J. Weaver, Inorg. Chem., 22, 1733 (1976). L. E. Bennett and J. L. Sheppard, J. Phys. Chem.,_§§, 1275 (1962). M. R. Hyde, R. Davies, and A. G. Sykes, J. Chem. Soc., Dalton, 1838 (1972). D. W. Carlyle and J. H. Espenson, J. Am. Chem. Soc., 29 2272 (1968). 9 c. Dultz and N. Sutin, J. Am. Chem. Soc., 22, 829 (1964). M. J. Burkhart and T. W. Newton, J. Phys. Chem., 73, 1741 (1969). '_— 235 160. A. Adin and A. G. Sykes, J. Chem. Soc. (A), 1230 (1966). 161. A. Ekstrom, A. B. McLaren, and L. E. Smythe, Inorg. Chem., 15, 1035 (1975). 162. C. A. Jacks and L. E. Bennett, Inorg. Chem., 22, 2035 (1974). 163. M. Faraggi and A. Feder, Inorg. Chem., 22, 236 (1973). 164. J. F. Endicott and H. Taube, J. Am. Chem. Soc., 22, 1686 (1964). 165. C. Lavallee, D. K. Lavallee, and E. A. Deutsch, Inorg. Chem. ~ .21, 2217 (1978). 166. T. J. Meyer and H. Taube, Inorg. Chem., 1, 2369 (1968), 167. T. J. Przystas and N. Sutin, J. Am. Chem. Soc.,_22, 5545 (1973). 168. J. P. Candlin, J. Halpern, and D. L. Trimm, J. Am. Chem. Soc. ‘22, 1019 (1964). 169. R. T. Wang and J. H. Espenson, J. Am. Chem. Soc., 22, 380 (1971). 170. R. J. Christenson, J. H. Espenson, and A. B. Batcher, Inorg. Chem., 22, 564 (1973). 171. R. Davies, M. Green, and A. G. sykes, J. Chem. Soc., Dalton, 1171 (1972). 172. J. N. Braddock and T. J. Meyer, J. Am. Chem. Soc.,_2§, 3158 (1973). 173. H. Fischer, G. M. Tom, and H. Taube, J. Am. Chem. Soc., 22, 5512 (1976). 174. T. W. Newton and F. E. Baker, Adv. Chem. Ser., 1;, 268 (1967). 175. M. J. Weaver, J. Phys. Chem., 22, 1748 1979). 176. P. D. Tyma and M. J. Weaver, J. Electroanal. Chem., in press. 177. M. J. Weaver and F. C. Anson, J. Phys. Chem., 22, 1861 (1976). 178. W. R. King and C. S. Garner, J. Phys. Chem., 22, 29 (1954). 179. J. H. Espenson, personal communication to M. J. Weaver, 1978. 180. Reference 82, p. 468. 181. 182. 183 O 184. 185. 186. 187. 188. 189. 190. 191. 192. 193. 194. 195. 196. 197. 198. 236 J. H. Espenson, K. Shaw, and O. J. Parker, J. Am. Chem. Soc., 22, 5730 (1967). P. Dodd and H. Taube, Z. Physik. Chem. (Frankfurt), 22, 92 (1965). B. R. Baker, M. Orhanovic, and N. Sutin, J. Am. Chem. Soc.,_8_2, 722 (1967). J. C. Hindman, J. C. Sullivan and D. Cohen, J. Am. Chem. Soc., .22, 2316 (1959). J. C. Sullivan and J. C. Hindman, J. Phys. Chem., 22, 1332 (1959). J. Podlaha and J. Podlahovfi, Coll. Czech. Chem. Commun., 22 3164 (1964) . G. Jones and W. A. Ray, J. Am. Chem. Soc., 29, 1571 (1944). L. Meites, J. Am. Chem. Soc., 12, 2476 (1953). B. Metz, D. Moras, and R. Weiss, Acta Cryst, 329, 1377 (1973). W. M. Latimer, "The Oxidation States of the Elements and Their Potentials in Aqueous Solutions", Second Edition, Prentice-Hall, New York, 1952, p. 185. M. C. Day and J. Selbin, "Theoretical Inorganic Chemistry", Second Edition, Reinhold, New York, 1969, p. 114. Reference 38, p. 507. I. M. Kolthoff and J. J. Lingane, "Polarography", Second Edition, Interscience, New York, 1952, Chapter XI. C. Luca, V. Magearu, and G. Popa, J. Electroanal. Chem., 22, 45 (1966). G. Popa and V. Magearu, J. Electroanal. Chem., 22, 85 (1969). R. P. Buck, J. Electroanal. Chem., 2, 295 (1963). M. E. Macovschi, J. Electroanal. Chem., 22, 457 (1968). E. Mei, L. Liu, J. L. Dye, and A. I. Popov, J. Solution Chem.,_§, 771 (1977).