MSU LIBRARIES “ RETURNING MATERIALS: P1ace in book drop to remove this checkout from your record. FINES wi11 be charged if book is returned after the date stamped below. 258 CHAPTER V INFLUENCES 0F SPECIFIC REACTANT-SOLVENT INTERACTIONS ON ELECTRON-TRANSFER KINETICS AND THERMODYNAMICS A. The Influence 9; Specific Reactant-Solvent Interactions 93 Intrinsic Activation Entrogies for Outer-Sphere Electron Transfer m (Accepted for publication in J. Phys. Chem.) 1- W In recent years increasingly detailed and sophisticated theories of outer-sphere electron-transfer kinetics have been formulated.56 These enable rates and activation parameters to be calculated from reaction thermodynamics together with reactant and solvent structural information. Although treatments of inner-shell (intramolecular reactant) reorganization have reached a high degree of sophisti- 23 cation, the important contribution to the free energy barrier arising from outer-shell (noncoordinated solvent) reorganization is usually treated in terms of the classical dielectric continuum model as 13 originally formulated by Marcus. While comparisons between theory and experiment for bimolecular outer-sphere processes show reasonable agreement in a number of cases,10 significant and often large 259 discrepancies still remain (cf. Chapter VII). Among other things, such discrepancies call into question the quantitative validity of the dielectric continuum model, especially in view of the well-known failure of similar treatments to describe the thermodynamics of ion salvation. In principle, a useful way of monitoring the influence of outer- shell solvation upon electron-transfer energetics is to evaluate en- tropic parameters since these are expected to arise chiefly from the changes in the degree of solvent polarization associated with electron transfer. The activation entropy,AS*, as for other reorganization parameters, can usefully be divided into "intrinsic" and "thermodynam ic" factors: 245 ’2“ AS 'AS. + aAS m (5.1) where the coefficient a is predicted usually to be close to 0.5.247 The intrinsic activation entropy, ASE“, is that component of [38* that remains in the absence of the entropic driving force A80. When esti- mating values of 138* from Equation 5.1, it is usual to employ experi- mental values of A80 and yet also employ values of ASL"; calculated from dielectric continuum theory. Although these calculated values of A82“ are often small, given that the values of [33° are often much larger and more variable than those calculated from the dielectric continuum model, it is reasonable to inquire if a more trustworthy :- method for estimating A8 int could be formulated. 260 A useful and often enlightening approach for understanding electron transfer processes both on a conceptual and an experimental basis is to examine the thermodynamics and kinetics of electrochemical reactions:53-55’248 Ox + e'(¢m) Red (1.12) where o!“ is the (Galvani) electrode-solution potential difference. Although absolute values of on cannot be evaluated with useful accuracy. the temperature dependence of ¢m can be obtained using a nonisothermal cell arrangement?“55 This enables the entropic change induced by reduction of a single redox center. the so-called “reaction entropy" Asgc, to be determined from the temperature dependence of the 55 standard (or formal) potential ¢: under these conditions. Activation parameters for such electrochemical "half reactions" can be obtained using an analogous procedure.53‘54 These quantities provide insights into the structural changes accompanying electron transfer at each redox center that remain hidden for homogeneous bimolecular reactions. The aim of this section is first to provide a simple physical picture, based on electrochemical half-reactions, of the origin of the intrinsic activation entropy in homogeneous and electrochemical redox reactions. With this background a new approach for estimating As:nt will be outlined based on reaction entropy data whereby the effects of specific reactant-solvent interactions can be taken into account. De- spite their potential importance, such interactions have yet to be con- sidered even in the more sophisticated theories of electron transfer. 261 2- mgmmmm The actual entropic barriers, A3: and As:, to electron transfer for the forward (reduction) and reverse (oxidation) electrochemical reactions at a given electrode potential have been termed "ideal" 53 ,54,248 activation entropies. These can be formulated as * * o A a A A 8f a 3re + sint,e (5'23) * s: I — A o A . ASr (a 1) Sn + Sint,e (5.2b) * where a is the electrochemical transfer coefficient and Asint e is the D so-called ”real" (or intrinsic) electrochemical activation entropy, i.e. that which remains after accounting for the entropic driving force.53 For convenience, we shall assme that the interactions between the reactant and electrode, and between the reactant pair in homogeneous solution, are weak and nonspecific (i.e. the "weak 249 * Under these circumstances AS. is related int,e to the intrinsic activation entropy for the corresponding self-exchange interact ion" 1 imit) . reaction by (5.3) Relationships such as Equation 5.3 reflect the fact that homogeneous outer-sphere reactions can be regarded as coupled reductive and oxidative electrochemical reactions. 262 Equations 5.2a and 5.2b point to a key difference between homo- geneous self-exchange and electrochemical exchange reactions: the latter are characterized by a net entropy driving force A8:c even when the free energy driving force is zero. This results from the inherent chemical asymmetry of the electrochemical half reactions. This entropy driving force Contributes to the forward or reverse entropic barrier for each redox center to an extent determined by the difference in (hypothetical) charge between the oxidized or reduced reactant and the transition state, namely aor (o-l). The transition state of course never acquires a fractional charge since electron transfer occurs approximately independently of nuclear motion, but nonetheless is char- acterized by a polarized solvent environment appropriate to a molecule possessing such a charge.13’56 According to the theoretical approach of Marcus,13 solvent reor- ganization to form the transition state can be viewed as occurring by a hypothetical two-step process. First the charge of the reactant is slowly adjusted to a fractional charge approximately midway between the reactant and product charges, with attendant reorientation of the sur- rounding solvent.250‘251 Then in a rapid step (much faster than solvent motion) the transitionrstate charge is reset to that of the reactant. Taken together, the energies of the two steps are equivalent to the nonequilibrium.solvent polarization energy. 0n the basis of the conventional dielectric continuum approach, the energetics of the first step are determined by the static solvent dielectric constant es, while the optical (i.e. infinite frequency) dielectric constant e 0? determines the energy of the fast second step. We shall term these two 263 steps the “static” and "optical" components, respectively. Generally the optical component is anticipated to provide the dominant contri- bution to the free energy of solvent reorganization due to the relative magnitudes of 80p and 5'. However, the temperature coefficients of the two dielectric constants are such that in many solvents the optical and static components are calculated to contribute roughly equally to the entropic component of the solvent barrier. The conventional calculation of the solvent reorganization energetics involves an application of the Born ionrsolvation model to 250.251 transition state theory. The Born model predicts that entropies 252 It is of ions will vary with the square of the charge number. reasonable to suppose that the static component of the electrochemical transition-state entropy will also depend on the square of the effective charge. The differences in static entropy between the transition and ground states should be appropriately weighted fractions of the total entropy difference Asgc between the two ground redox states. We can therefore express the static components of the forward and reverse electrochemical activation entropies as Aspmm) - {[(n + 1)2_(n + 1 -a)2]/[(n + nz-uznAsgc (5.4.) A8:(static) ' {In2 - (n + l -o)2]/[(n + l)2 - n21}As:c (5.4b) where n and n+1 are the charge numbers of the two forms of the redox couple, and (n+l-a) is the effective transitionestate charge. 264 It can be seen from Equations 5.4a and 5.4b that even for a transition state that is symmetrical with respect to charge, i.e.cz- 0.5. that A8;(static) will differ from -ZB:(static). In other words, the transition state will not lie midway in terms of entropy between the reduced and oxidized states even though it may be equally accessible in terms of free energy from either oxidation state. This mismatch of the energetics of the forward and reverse half reactions follows from the liggg; variation of driving force contributions with charge (Equation 5.2), coupled with the quadratic dependence of static entropy on charge. Equations 5.2a and 5.2b can be combined to yield * int,e AS - (1-0)AS; +aA S: (5.5) The intrinsic activation entropy therefore is a measure of the extent of the mismatch between forward and reverse half-reaction entropic barriers after normalizing for driving force contributions. This is * f a . . and A8r exactly cancel. The connections between the various entropic seen most clearly whena.- 0.5 and the driving force components of AS quantities are illustrated schematically in Figure 5.1. The magnitude * of Asint is given by the vertical displacement of the curve AB from the chord (dashed line) to this curve. The curve AB describes the depen- * dence of ASin upon the effective ionic charge. 1: Equations 5.4a and 5.4b can be combined with Equation 5.5 to yield an expression for the static component of As:nt e: 5 265 158* int,e(static) - [o(l—o)/(2n+1)]AS:c (5.6) Taking a . 0.5 and inserting the Bornian expression for the reaction entropy:253 As:c -(Ne2/2r€§)(d€sldT) [(n+l)2-n2] (5.7) into Equation 5.6 yields: * A3 int,e(static) - -(Ne2/8r€§) (dealdT) (5.8) where N is the Avogadro number, e is the electronic charge, and r is the reactant radius. Note that the apparent dependence of Asint,e (static) on reactant charge in Equation 5.6 has now been eliminated. Equation 5.7 can be compared with the relation obtained from the temperature derivative of the usual dielectric continuum.expression for the reorganization free energy:13’54 * AS int,e ' [(Ne2/8)(1/r-l/m] [(l/eip)(deop/dT)-(l/c§) (desldTH (5.9) Equations 5.8 and 5.9 differ in that the latter takes account of image stabilization of the ion in the vicinity of the electrode by including the ionrimage separation distance R; furthermore, the optical portion of the activation entropy is included. This term is similar in form to the static term since it is assumed, based on a linear response of solvent polarization to the field of the ion, that the optical portion 266 g, A \ \\ A53... , AS‘,‘ 3 g. (a-MS}. \\ o .3 A55, \ LIJ \\ . .53 \ 1531 c \ .9. 0A5}. \ \\ \\ $3, 5 s l l l n m l-a ml Effective Ionic Charge Figure 5.1. Schematic representation of the ionic entropy of an individual redox center as a function of its effective ionic charge during the electron transfer step. See text for details. 267 Of As:nt e also varies with the square of the effective charge of the 9 transition state.56b Similarly. from Equations 5.3 and 5.6 the static portion of the intrinsic entropy for homogeneous self-exchange reactions can be expressed as ASInthtatic) ' [205(1-o)/(2n+1)] As:c (5.10) Again, for <1 - 0.5 and on the basis of the Bornian model (Equation 5.7), this leads to * . 2 2 Asint(static) - (Ne l4r88) (dEBIdT)] (5.11) This is identical in form to the dielectric continuum expression for A * ( f ' . 8int c . Equation 5.9). As;t - (Ne2/4) ($.- é) ((1/efip) (deep/dT)-(1/€:)(d€8/dT)] (5.12) allowing again for the addition of the optical term and the presence of the nearby coreactant through the internuclear distance term R. 3. 3&5; Chemical Environments. Incorpogatigg Specific Egactant-Solvent Interactions ig.Activation Entropy Calculations, In general, the experimental values of Asgc differ widely from the continuum predictions of Equation 5.7. In water, for example as: for the Cr(1120)2+/2+ couple is seven times greater than predicted, 268 while the experimental value of AS:c for F'e(bpy)g+/2+ is less than a third of the theoretical value. Furthermore, the expected variation of reaction entropies with solvent dielectric properties is not observed.253-255 The discrepancies between theory and experiment have variously been attributed to dielectric saturation, hydrogen bonding between reactants and solvent,55 long range solvent structuring,253-256 and hydrophobic interactions.257 Consequently, in view of Equation 5.10 dielectric continuum theories of solvent reorganization are not expected to provide accurate estimates of intrinsic activation entropies. Nevertheless Equation 5.10 suggests a means of incorporating the numerous factors neglected in the dielectric continuum treatments. Rather than employing estimates of As:c based on Equation 5.8, experi- mental values of AS:c can be used to determine the static component of * A81“. Therefore instead of Equation 5.12 the intrinsic activation entropy can be expressed as * _, 2 1 1 2 As,1m (Re /4) ‘F" E) [(1/eop) (deop/dTH + Asgc/(an + 2) (5.13) * int’ Equation 5.13, is unchanged from Equation 5.14; however the static The optical component of AS the first term on the r.h.s. of component embodied in the second term is taken instead from Equation 5.10 with a - 0.5 (As expected, a is commonly observed to be close to 0.5 for outer-sphere electrochemical reactions). Equation 5.13 is therefore anticipated to yield more reliable values of Asint’ at least in the weak-interaction limit, since it circumvents the known severe 269 limitations of the Born model for calculating static entropies. The latter model is retained for estimating the optical component .1; lieu of any direct experimental information to the contrary. The justification for this approach is that the Born model is likely to be much more reliable for estimating the optical rather than the static component in view of the relative insensitivity of 80p to solvent structure. Thus the extensive local perturbations in solvent structure induced around an ionic solute that are responsible for the failure of the dielectric continuum model for predicting ionic solvation thermo- dynamics should have a much smaller influence on the intramolecular electronic perturbations which constitute the optical component of the reorganization barrier. A comparison between values of As:at calculated from Equations 5.12 and 5.13 for some representative redox couples in aqueous media is presented in Table 5.1. Whereas the dielectric continuum model * (Equation 5.12) predicts that Asia will be small and nearly inde- t pendent of the chemical nature of the redox couple, somewhat larger and a more varying values of 13in are predicted by Equation 5.13 since this t takes into account specific reactant-solvent interactions via inclusion of the experimental values of Aszc. Further, the latter relation * predicts markedly larger variations in ASint with solvent than obtained with the former relation, resulting from the much greater sensitivity 0f Asgc to the solvent than predicted by the Born model.253-255 Although the differences between Equations 5.12 and 5.13 have been emphasized here, it should be noted that the As:nt values obtained by the latter are still relatively small. An interesting result is 270 that for multicharged reactants very large thermodynamic solvation effects translate to much smaller intrinsic entropic barriers. For example, the Fe(HZO)Z+/2+ self-exchange reaction involves thermodynamic l l l entropy changes amounting to 360 J deg- mol-1 (180 J deg- mol- for each half reaction) which yields an entropic contribution of just 13 J -l 1 deg mol- to the Franck-Condon barrier (Table 5.1). Still, the effects are large enough to warrant consideration. For example, for 3+/2+ * the “(320)6 reaction ASin should contribute a factor of five to t the self-exchange rate constant at room temperature. This effect is therefore comparable in magnitude to the nuclear tunneling corrections and nonadiabatic electron tunneling factors which have been emphasized in the recent literature.23’258 4. Copparisons with Experipgnt In addition to the calculated values of Asint’ some “experimental" values for these homogeneous self-exchange reactions, * Asint(°xp)’ are given in Table 5.1. The latter were extracted from the measured activation enthalpies, Ant, and the rate constants, k, using * I - * k lipl‘nvn exp(ASint/R) exp( AH IRT) (5.14) where KP is the equilibrium constant for forming the precursor complex immediately prior to the electron-transfer step, rn is the nuclear tunneling factor. and x”: is the nuclear frequency factor.9’23’258 [Note that the activation entropy in Equation 5.14 can be directly . . . . * identified with A31“ since A30 - 0 for self-exchange reactions 271 (Equation 5.1)]. The values of RP and Tn were calculated as described in Chapter VII and reference 10. The values of As:nt(exp) were corrected for the variation of Pa with temperature by calculating this quantity using the relationships given in reference 23. It is seen in Table 5.1 that the values of Asznt(exp) are uni- formly smaller, i.e. more negative, than the estimates of ASEfit from both Equation 5.12 and Equation 5.13. Such negative values of As;nt (exp) are commonly observed for homogeneous outer-sphere reac- tions. They have been variously attributed to an unfavorable contri- bution to the precursor work term arising from reactant-solvent inter- actions, to the occurrence of nonadiabatic pathways, and to steric factors.“’207 In any case, in view of the present discussion it ap- pears likely that these negative values of Asint reflect properties of the bimolecular precursor complex rather than those of the individual redox couples; i.e. reflect the modification of the solvation environ- ment around each redox center brought about by its proximity to the co- reactant necessary for electron transfer. In fact, the inclusion of specific reactant-solvent interactions in the calculation of As:nt for the weak interaction limit by employing Equation 5.13 rather than Equa- tion 5.12 leads in most cases to more positive values of As:nt (Table 5.1). Before accepting this conclusion, however, it is worth examining further the various assumptions embedded in Equation 5.13. Given the breakdowns observed thus far in the Born solvation model, it is possible that the assumed quadratic variation of entropy with charge is also incorrect. The magnitude of the intrinsic activation entropy 272 Table 5.1. Intrinsic Activation Entropiem for Selected Homogeneous Self-Exchange Reactions, ASLc (J deg-1 mol-1). calculated without (Equation 5.12) and with (Equation 5.13) Consideration of Specific Esactant-Solvent Interactions. and Comparison with Experiment. a b a d Eedox Cou 1e Solvent ' r R A 3* A s' 35* P ' ' int int int Equation 5.12 Equation 5.13 (experiment) 3+/2+ Fe(0!lz)6 a 820 3.3 - -1.5 ‘ 13 -62 3+/2+ Hafiz)6 320 3.3 -l.5 12.5 -61 Eu(NE3):+/2+ s20 3.3 -1.5 2.5 -24‘ Co(ed§+’z*f 320 «.2 -1.5 9.5 -as 9 many)?”2+ 320 6.7 -1.0 -4.5 -- ferricinium- 820 3.8 -l.5 ~15 -— ferrocene ferricinium- methanol 3.8 -6.5 -9.5 -64 ferrocene ferricinium- nitromethane 3.8 -6.0 16 -29 ferrocene ' akeactant radius, used to calculate ASInt (Equation 5.12). Values taken from references 10 and 255. bIntrinsic activation entropy, calculated from Equation 5.12 using the listed values of r and assuming that 'Zr (Reference “.82)- Literature values of t ,(dc Id'I'). c . (dc /d‘l‘): water -e -78.3. (deg/d )--0.365, :0 -1.7s, (dc /dT)--0.00024 (valuessfrom retereeEB 165,°Bp. E61, 3224); Serhenei- r -32.6, (dc /dT)--B.20 (reterSBcee 263); c -1.76. (dt ldT)--0.0011 (reference 260, p. 145); n tromethanegc.-35.5, (dc./dT)--O.l6 (rererSBee 265); cop-1.90, (drop/dr)--o.ooio (reference 266, p. 391). c Intrinsic activation entropy, calculated similarly to footnote b above, using the experimental values of 68:: taken from references 55 and 255. dValues extracted from published rate dattzby_!sipg Equation 5.16 and correcting for nuclear tunneling effects. Values of K v are circa. 3 x 10 H 5 (see reference 10 and Chapter VII). Literature sources for rate datg:n Fe(l!20)63+/2+ - reference 266, V(820)g+/2+ - reference 267, Ru(NHg)a+/2+ - reference 268,Co(en) 3442* - reference 269, ferricinium/ferrocene-reference 270. ‘Brown and Sutin (reference 28) have questioned the accuracy of this result. based on the more negative AS‘ value for the au(qu+ - Eu(Nfl3)§ cross reaction. fen - ethylenediammine gbpy-2,2'-hipyridine 273 obtained from Equations 5.13 is closely connected to the functional dependence of entropy on charge. For example a linear dependence leads to a value of zero for Asint' Other functions might lead to large imbalances of forward and reverse entropic barriers and therefore substantial intrinsic activation entropies. Since most couples exhibit positive values of As:c a fractional dependence of entropy on charge a would normally be required to deduce negative values of ASin The t' entropy-charge relation was the subject of a number of detailed examinations and some controversy in previous years, and apparently was never unambiguously resolved.259"261 One reason for this was the difficulty of varying the ionic charge while holding constant the other relevant parameters such as ionic size, ligand composition, coordination number. etc. In order to determine the relation between entropy and charge for a prototype system we examined the reaction entropies of ruthenium tris bipyridine, for which oxidation states 0, I, II, and III are accessible in acetonitrile. The experimental details are given in reference 253. By employing the same compound in various oxidation states the numerous complications and ambiguities inevitably involved in previous studies are avoided. The reaction. entropies thus obtained for the Eu(bpy)g+/2+. Ru(bpy)§+/+. and Eu(bpy);/o couples, respectively, in acetonitrile (containing 0.l‘l_l_KPF6 supporting electrolyte) are 117, 71 1 mol-l. If these data are recast as relative single ion and 23 J deg- entropies, 8° + K, where K is an unknown constant quantity, a straight- forward variation of entropy with the square of ionic charge is evident (Figure 5.2), supporting the validity of Equation 5.13. 274 It is possible of course that the ruthenium trisbipyridine reactions represent an atypical case. (However, additional evidence is assembled in Section V. C). Another way of exploring the possibility that the negative experimental values of A8;t might arise in part from mismatches in the thermodynamic entropic changes occurring in each half reaction is to examine if the magnitude of A8:nt depends on the sum of the constituent As:c values. The larger these entropy changes, the a a larger should be the mismatch in ASf and ASr for each half reaction, * yielding larger (or more negative) values of ASin However. such an t. examination for about thirty self-exchange and cross reactions shows no signs of such a systematic trend. (Details are given in Chapter VII). * In addition, the experimental values of ASin also show no discernable t dependence on the magnitude of the reorganization barrier. comparable . * . . . negative values of AS. being obtained even for extremely rapid int 207 reactions. This provides evidence that these negative values are associated either with an entropically unfavorable work term and/or nonadiabaticity, rather than residing in the elementary reorganization barrier to electron transfer of which the estimates of As:gt obtained from Equation 5.13 form a part. Nevertheless, the method of calculating Asint embodied in Equation 5.13 is considered to be useful since it provides a reliable * . . estimate of ASin for the limiting "weak interaction" case where the t solvating environments of the two reactants do not modify each other, while accounting properly for the influence of the actual reactant- solvent interactions upon the entropic reorganization barrier for these isolated redox environments. It therefore provides a more trustworthy 275 p l l l l o 4 s (Ionic Charge)2 Figure 5.2. Plot of relative ionic entropy of Ru(bpy); (bpy- 2,2'-bipyridine) in acetonitrile as a function of the square of the ionic charge n. The solid line is the best fit line through the experimental points; the dashed line is the slope of this plot predicted by the Born model. 276 means of gauging the extent of influence of reactant-reactant inter- actions upon the activation entropy than is obtained by employing the conventional relationship (Equation 5.12), as well as supplying useful insight into the physical and chemical factors that determine this quantity. A related approach to that described here can also be employed to estimate the effects of isolated reactant-solvent hydrogen bonding on the intrinsic enthalpic component of the Franck-Condon barrier. This involves examining the solvent dependence of the half-cell redox 262 thermodynamics. Preliminary results indicate that such enthalpic effects are markedly larger than the corresponding entropic factors examined here, contributing several kJ mol-1 to the intrinsic free energy barriers for a number of reactions.262 These findings suggest that such specific reactant-solvent interactions may indeed account in part for the common observation that the experimental rate constants for homogeneous outer-sphere reactions are significantly smaller than the theoretical predictions where the outer-shell reorganization energy is calculated using the conventional dielectric continuum model.10 B. Utility pf Surface Reaction Entropies £2; Examining Eeactant-Solvent Interactions pp Electrochemical Interfaces, FerriciniumrFerrocene Attached to Platinum Electrodes (Accepted for publication in J. Electrochem. Soc.) 1.111531111122122 Reactant-solvent interactions are of prime importance to both the kinetics and thermodynamics of electrode processes. Since electro- 277 chemical reactions inevitably occur within the interfacial region, it is desirable to gain information on the nature of reactant salvation at the electrode surface as well as in bulk solution. Weaver. et al. have demonstrated that useful information on the latter for simple redox couples can be obtained from the so-called "reaction entropy", A836, determined from the temperature dependence of the formal potential, Ef, using a nanisathermal cell arrangement:55 As° - Nerf/er) (1 21) re ni ° Since the temperature dependence of the thermal liquid junction poten- tial in such a cell can be arranged to be negligibly small, A8:c essentially equals the difference, (3° 0 . . red 80:), between the ionic entropies of the reduced and oxidized forms of the redox couple in the bulk solution. The reaction entropies of simple transitionrmetal redox couples have been found to be extremely sensitive to the chemical nature of the coordinated ligands and the surrounding solvent, illus- trating the importance of specific ligand-solvent interactions to the overall redox thermodynamics.””84’2”-255 It would clearly be desirable additionally to determine reaction entropies for redox couples residing in the interfacial region. Such 0 "surface reaction entropies", Asrc s 9 would yield insight into the salvation changes accompanying the elementary electronrtransfer step for the redox couple in a particular interfacial environment. For redox couples present at sufficiently high concentrations at the inter- face to enable the formal potential for the interfacial (adsorbed) 278 redox couple, E2, to be measured, values of A8:c s can be obtained 8 directly from (cf. Equation 1.21): o f asrc’s Foum m «a muHHHAHoamoummuuH .onmaoo mo muHHHnnHomaH ou mam momsHooum mandamusmmoZm .moumonnH omaumuo HH Baum m.ou mumm mama moo oHuuooHoHv wnHma NH. m nOHummvm Baum voumHmUHmu .Homoam neon ecu 80mm vouonmnm own no om HmHunouoa Hmauom .omw monouowmu Somme a x on no nNN em II 11 No occuoo< an.o~ we oeH mm mm mm mOH oHHmanOuoo< w «H Hm oeH mm em mm (1 ocmHOMHam e 3 fl 8? .2. N2 2 m2 3263522353 & mm MH «RH NH: we H 1 HHH Hocmcuoz mn.0H H~1 m o HON on: «NH noun: Guam m.ou on w on on m.omm< mn9< umo>Hom w A omqv owe. Q m om< Q m o a onooomuow onHuHsHouna . onooouumw u loclocuoaonooouquIn monomuumloumwuam monmou mnooouuomlanHmHoHuuom GOHuaHomianm mom monomuu<1oommuam Mom HHIHoa H mom .HV monomunm nOHuomom one A>av mHmHunmuom Hmauom .~.m oHemH 285 the relatively electronerich ferrocene redox center to a greater extent than for ferricinium, leaving the former more difficult to oxidize and thereby yielding a positive shift in the formal potential. The syste- matic differences in reaction entropies seen between ferrocene and the derivatized couples can also be rationalized on this basis. Whether the differences in formal potentials between the adsorbed couple and its solution analog result from surface attachment or from differences in substitutent properties is not entirely clear. A curious aspect of the results is the marked solvent dependence of both the Asgc s and Asgc values. The magnitude of these quantities expected from purely continuum electrostatic considerations is given by Equation 5.17:255 Mano/133:) - -(e2N/2r€T)(dln€/dlnT) (5.17) where e is the electronic charge, N is Avogadra’s number, s is the dielectric constant of the solvent, and r is the radius of the ferricinium cation. The A8:c a values listed in the last column of 8 Table 5.2 are obtained from Equation 5.17. using literature values of e 274 and assuming that r - 3.88. There is clearly no general pattern of agreement between the experimental and these calculated quantities, the Born treatment predicting a much milder solvent dependence of asgc s 9 than is observed. Similar breakdowns of the dielectric continuum model in predicting reaction entropies have been found for several bulk solution couples in a number of solvents.55’84’253'255 286 A probable reason for the failure of Equation 5.17 is that the major property determining the entropy Of charge-induced solvent reorientation is the degree of "internal order" of the solvent (i.e., self-association and long range structuring induced by hydrogen bonding), rather than the macroscopic dielectric properties.275’276 Thus, a solvent having a high degree of internal order would be relatively unperturbed by a charged molecule, whereas considerable solvent ordering around the ion would occur in a medium having little intermolecular structure. Since such charge-dipole interactions will be absent for neutral ferrocene, a positive contribution to the reaction entrepy (S:ed - 83“) would be anticipated for the present redox couples, especially in relatively nonassaciated solvents. Criss and co-workers have suggested estimating the degree of internal order of a solvent from the difference in boiling point, Apr, compared to that for a structurally analogous hydrocarbon. These values of Apr are also listed in Table 5.2. Indeed, the Asgc,s values for the surface-attached ferrocene couple do for the most part vary as expected with the corresponding values of AT p' An unusual result which merits comment is the large negative 1 value of A8:c s (-50 J deg- mol-l) found in water (Table 5.2). A 8 small negative value of A8:c has previously been oberved for the bulk- solution ferrocene couple, also in water.255 This indicates that the net solvent ordering in the vicinity of the surface-attached redox center is less extensive in the cationic than in the neutral state, in qualitative disagreement with the expectations from an electrostatic treatment. These negative entropy values possibly result from donor- 287 .uoumn .o— “Homecuoa .o “ooHEmBLOu .o “ocHemauoUHaguoE1z .m nonsense (anuH: .o momeOuHomecuuaHv .m moHHHUH:Ououn .e mouncocmmu ocoHAQOma .n mom—Em5u0uH>:uoeHm .N "accumuo ._ "mu:u>H0m cu aux .ochOHuou :oHumHomanmc "moHomHo mono "Ac.n omamumv acouomuou momma Im>Huom coHusHomuxHam "moHucmHuu coao “an.n ouomHmv ocooonuou cozomuumuoumenzm "moncmHuu mmHHHm .HHN oucmnouon aouu cmxmu .uocaaz mauaooo< uco>Hom momuo> muco>Hom maoqnm> cu monaaoo unoccuuom (ancHoHunou mmocaucoHusHom mam mocomuumuoomuusm Haw mouaouuco :oHuommu mo uon .o.m shaman as}; Shnwu u< hzmgm Om On 0? On ON 0. _ . _ _ i l 0 e r b o O 1. o I_IOW..°3P-P/AdOHlN3 NOIlDVBH ' O. r as l O Q' 288 acceptor interactions between the cyclopentadienyl rings and the acidic water hydrogens.255 (An alternate and additional explanation in terms of an entropy change associated with "solvent disruption" is offered in Section V. C). Since the electron density on the cyclopentadienyl rings will be greater in the reduced state, such specific solvent interactions should be enhanced leading to increased solvent ordering and a decrease in entropy compared with that for the oxidized state. If such an explanation were correct a correlation between A8:C,B and the acidity of the solvent might be expected. Figure 5.6 shows a plot of the reaction entropies for the adsorbed couple and its solution analog as well as for unsubstituted ferrocene versus the solvent "acceptor number“ which is an empirical measure of the electron- accepting capabilities of the solvent.277 A reasonable correlation is indeed observed. Apparently, both noanornian ianrdipole interactions and specific donor-acceptor interactions might be important in determining the redox properties of the surface-attached as well as solution ferrocene couples. The present work demonstrates the feasibility of determining surface reaction entropies and illustrates the utility of these measurements for elucidating the various elements of interfacial reactant-solvent interactions. Given the sensitivity of Asgc,8 measurements to the solvent structure it is suggested that this approach might also usefully be employed to gain insight into reactant salvation in polymer film electrodes for which the question of solvent penetration within the film is of current interest. 289 C. Size, Charge, Solvent and Ligand Effects pp Reaction Entropies 1. Introduction Relative entropies of redox reactions were widely measured in the 1950’s and 19602s in order to examine basic notions concerning ionic 281-287 salvation. Interest in this topic was revived in 1979 with the report by Weaver and co-workers that absolute measures of the entropy 0 0 A 0 - - 0 m a difference 31“: ( sred sax) between the two ions forming a redox couple could readily be obtained from nonisothermal electrochemical experiments.55 Numerous papers on reaction entropies have appeared since then.84,253-255,271.28l-304 These have tended to emphasize either the value of such measurements in unraveling the details of solvent reorganization in connection with electron transfer 55.84,253-255.293,294,297,304 dynamics or their usefulness for gaining . . . . . 300-302 . information concerning the salvation of metalloproteins, peptide complexeszgs’299 and other biological model compounds.303 Although many insights have been gained into both problems, a number of puzzles remain. This section describes some empirical relations which have been uncovered concerning the dependence of As:c values for simple tran- sition metal couples on reactant size and charge and an the nature of the solvent, and seeks to provide interpretations at the molecular level. Besides offering predictive power. it is suggested that these correlations and interpretations can rationalize some of the more curious findings of earlier studies. 290 Z-EmLLta The data examined here have been taken largely from previous reportsss’84’89’253-255 although a fair number of new results are included. (The new results are summarized in Appendix I). In each case the redox couple is substitutionally inert in both oxidation states. Values of Asgc in. water, dimethylsulfoxide (DNSO) and acetonitrile for several "3+/2+" couples are plotted against reactant radius in Figure 5.7. For spherically nonsymmetrical complexes (e.g. 3+/2+) Ru(NH3)5py , the effective radius is taken as equal to half of the cube root of the product of the diameters along the three ligand-metal- ligand axes.28 Excluded from the comparisons at this point are redox couples where a difference in spin state occurs between the oxidized and reduced forms. It is evident that there is a good linear fit of the data in water. with the exception of the results for three couples containing aqua ligands. (These three could not be examined in the other solvents). In acetonitrile and DHSO the reaction entropy also varies with -r, but apparently not in a strictly linear fashion. Somewhat better linear correlations, at least in nonaqueous solvents, are found, with l/r (Figure 5.8). Similar correlations were obtained in solvents other than the three for which data are shown, but these are omitted from the plots for clarity. The dependence of As:c on charge was examined by monitoring consecutive reduction reactions of ruthenium- and chromium tris bipyridine complexes. For both of these, at least four redox states are accessible in acetonitrile and acetone. Figure 5.9 shows that the 291 i zoo~ ‘ ‘ _ 12 (3 1225 7 "r fig C: 3 9A 4 E! E! ‘A 1: —- s 1 14 (a 8’ 8 4‘ 2 -:: 1()()r- ‘3 - '-5 12 b 8 ‘8 ‘2‘;_ 11 g; (a (D ()7 4 ‘°<>5(5 1 <3 223 ()- 3 L 1 1 1'7 13 '7 ES r,A Figure 5.7. Reaction entropy versus effective radius of reactant. Key to solvents: Key to reactants: (0) water; (A) dimethylsulfoxide; (D) acetonitrile. + (1) c:(bpy)33+; (2) Fe(bpy)33 ; (3) Ru(bPY)33+; (4) c-Ru(NH3)2(bPY)23+; (5) c-Ru(H20)2(bPY)23+; (6) c-Ru(n20)2(bpy),3+; 3+ (7) Ru(NH3)4bpy 3+ 3+ (11) RU(NH3)6 ; (12) 05(NH3)6 _: (13) Ru(NH3)5H20 3+ ; (8) Ru(NH3)aphen : <9) Ru33+i (1°) Ru(“3’5” 3+. 3+ 3+ (14) Ru(NH3)4(H20)2 ; (15) R“(H20)6 . 3+ 292 200 7 / /9/// T 4/’/ 95/] '5 ,0 A/ s- /D // 8 (u /’ '5 100 b / fl . a 8 /9/ (1) 7° / C) < /// 9 4 l’,/’ Figure 5.8. Reaction entropy versus l/r. Keys to solvents and reactants as in Figure 5.7. J (:leg“I mol‘I o I'C' AS 293 100 " 0| (3 l 2 2 zox' 2 red Figure 5.9. Reaction entropy for ruthenium tris bipyridine couples in acetonitrile versus the difference in the squares of the charges on the oxidized and reduced states. 294 reaction entr0py for Ru(b1>y)§"'un in acetonitrile varies with (lied -z§x), where 2°x and zred are the charges of the oxidized and reduced states of each couple. Unfortunately. there are only a few complexes and solvents for which consecutive electron transfer reactions can be examined. The connection between reaction entropies and the nature of the solvent is illustrated in Figures 5.10 and 5.11. Linear correlations are found between A8:c and the so-called solvent acceptor number.277 regardless of the reactant charge, size, electronic state or ligand composition. The As:c values for Ru(NH3)SN032+/+ are taken from the dissertation of Dr. Saeed Sahami,89 while those for Co(EFNE-axosar-H)2+/+ (structure in Figure 5.12) were measured by Dr. Peter Lay and this author. 3. Discussion The simplest theoretical treatment of reaction entropies is based on the Born electrostatic model in which the solvent is treated as a continuum.252 According to this model: a _ 2 2 _ 2 Asre (e Nl2e'l'r)(dlr1e/d'lf)(zox zred) (5.8) where e is the dielectric constant of the solvent. Although this approach has rightly been criticized for failing to provide accurate overall estimates of Asgc, the empirical analyses confirm the predicted variations with 1/r and (22x - 22 ). The chief problem seems to lie red in the representation of the solvent. 295 (J l l I 0 ' 20 40 60 Solvent Acceptor Number Figure 5.10. Reaction entrapy versus solvent acceptor number. Key to reactants: (I) Cr Rh, therefore “6:. < 2Ast’e. The quantity C in Equation 6.3b accounts for this inequality which is also the origin of the inequality sign in Equation 4.15; from Equations 6.3 and 6.4 308 2 e l l 1 l c - ( --— )(- - —) (6.5) '4 'Eh Re Ebp E:s Equation 4.15 can therefore be written in the more general form e I h - 2 108 (hex/19) log (kex/Ah) c/2.303 RT (6.6) For a series of reactants having a similar size and structure, as for the present aqua couples, Rb and Re and hence C should remain approximately constant. It is desirable to obtain a self-consistent set of experimental values of kg: or kg: for comparison with theoretical predictions obtained from calculated values of Aczx using Equations 6.2 and 6.3. This task is less straightforward than is commonly presumed for two reasons. Firstly, the experimental values of kzx or kzx may not refer 3+/2+ ) at least one of the “Q to outer-sphere pathways since (except for Ru aqua reaction partners is substitutionally labile so that more facile inner-sphere pathways may provide the dominant mechanism. Secondly, values 0f 1&2: derived using Equation 6.1 from rate constants for appropriate outer-sphere cross reactions rely not only on the availability of values of k:x for the coreacting redox couple along with values of K12, but depend also on the applicability of this relation.3m‘311 The resulting estimates of kg: for different redox couples are often difficult to compare since large systematic errors can be introduced by the use of cross-reaction data involving structurally different coreactants, inappropriate electrode potential data. etc. 309 In spite of their direct relationship to the desired intrinsic barriers, electrochemical exchange rate data have seldom been utilized for this purpose. One reason is that these data have commonly been gathered at ill-defined solid surfaces where the work terms arising from double-layer effects are large and unknown, precluding quan- titative intercomparison of the results. However, Weaver and coworkers have determined accurate electrochemical rate data for Vi+l2+, Cr3+/2+ 8Q Buzz/2+. Ruiz/2+ and Feizn+ at the mercury-aqueous interface under conditions where the work terms are small and can be estimated with 228’312‘313 The interactions between the reactant and the confidence. metal surface are likely to be weak and nonspecific, so that the electrode can be viewed as an inert electron source or sink that does not influence the electronrtransfer barrier. This allows information on the electronrtransfer barriers to be gathered for individual redox couples as a function of the thermodynamic driving force. Such information is largely inaccessible from the kinetics of homogeneous electron transfer. In the present section, suitable rate data for electrochemical and homogeneous reactions involving aqua redox couples are analyzed and compared using Equations 6.1 and 6.6 in order to ascertain as unambig- uously as possible how the kinetics of outer-sphere electron exchange depend on the metal redox center. 2. Rap; Constants for Electrochemical Exchapge Table 6.1 contains a summary of rate parameters for the electrochemical exchange of Roy/2+, V3+l2+, Fey/2+ u3+l2+. aq aq as a , E and .uxou ecu nH voumonnH mm mommaHummk .m.c xo nOHumaom mems me mom Ammmv ox aomm monHmuno .owcmcuxm HmuHamnoouuomHo now unnumnoo oumu mouooquUIxuozu .wNN mam HwH mooeououmu Baum mommaHumm no vmchuco mouhHouuooHo co>Hw mH m on HmamH mmswuHm mmouum HmHunmuom mumBonumm< .wmmvcmHmH moommummmu Baum mmnHmuco “HmHuwouoa Hmauom um muhHouuomHm moumum nH omouuoWHo museums um ummumcoo some «a osHm> Aucmumaamv mmusmmoZU .hu0umuoan chu Baum memo mocmHHnsmma me mmomoamHmmam momma monmuommu Baum :mxmu mosHm> .uomm um oquomuomHo mmumum nH onaoo Ramon mam HmHucmuom HmauomQ .+mAmmovz mo :oHumnouoummm ucmOHanme mmmuamam ou mHom Ame oHImAV ocmHonmam monHmunoo mumHouuomHmo n1om x o.H nae- naeoaovem m e.o c I am on I K I . H once N mm muoa N nee nee z e o +N\+m o o l om um I K I . 3 o muoa m cm euoe e mNe use z e o +~\+m u n on x am on x N wee ease m e.o were Ql Kml +N\+M on x H m- OH x m.e Nee- eeee m 3.0 ee> m1 m1 +~\+m on x Nu OH on x m oN- omen m e.o seem N- m- +~\+m Him So >8 Him So .m.o.m .m> >8 oumHomuomHm mHmaou Momma m no N0 .3 e Aeeeeee em a m o a u .UOmN um oomwuoueH maom=o<|>usouoz one on moHaaoo xommm ono< HHHV\AHHHVZ mean no mmmmcoxm HmoHaonoouuoon one new mmouoamumm moHammamoauoce voumHom mam mOHuonHM .H.o oHan 311 Orig/2+ at the mercury-aqueous interface, using potassium hexa- fluorophosphate and lanthanum perchlorate supporting electrolytes. These experimental conditions minimized the extent of the electrostatic double-layer effect upon the apparent rate constants for electro- chemical exchange k:x(app) (i.e. the "standard" rate constants measured at the formal potential Ef for the redox couple concerned), enabling values of the work-corrected rate constants kg: to be evaluated with confidence using314 e F ln kex ln ke‘(app) +RT(zr-acorr) 4) d (6.7) where z r is the reactant charge number. a corr is the work-corrected cathodic transfer coefficient, and ¢ (1 is the potential drop across the diffuse layer. Details of this procedure are given in references 312 and 313. The KPF6 electrolyte provides an especially suitable medium for this purpose. This is because (id is small over a wide potential range positive of the potential of zero charge (-440 mV vs. -s.c.e.) since the positive electronic charge density at the electrode is matched approximately by the charge density due to specifically adsorbed PF; anions.313’314 The rate data in Table 6.1 all refer to acid-independent pathways. In contrast to homogeneous reactions between aqua cations, the rates of most of these reactions are independent of pH at values (l.315 312 3+/2+ The formal potential for Feaq is too positive (495 mV vs. s.c.e. in 0.4 NLKPFé) to allow rate measurements in the vicinity of Ef to be made at mercury since anodic dissolution of the electrode occurs 3+ aq beyond about 375 mV. However. the electroreduction of Fe was found by Tyma313 to be sufficiently irreversible so that cathodic d.c. and normal pulse polarograms were obtained over the potential range 300 to 4 l 0 mV, yielding values of kapp in 0.4 §,KPF to 4 x 10- cm s- at 300 l 6 mV, and 0.1 cm s- at 0 mV vs. s.c.e. Extrapolation of the cathodic Tafel plots, (i.e. ln k: vs. E, where k: is the apparent cathodic PP PP rate constant), was therefore required in order to extract k:‘(apP). However. this procedure can be applied with confidence: the work- corrected cathodic transfer coefficients “cart for several other aqua couples are close to 0.50 (10.02) over a wide range of cathodic over- 313 3+ potentials. aq The observed transfer coefficientsapp (0.48) for Fe reduction in 0.4,§._KPF6 indicates that the potential dependence of the double-layer effects is likely to be small, as expected. Consequently. the resulting value of k:‘(app), 2x10.5 cm s-l, is likely to be within a factor of 2- to 5-fold of kg‘; we have therefore set an upper limit -1 of lxlO-4 cm s for kg: in Table 6.1. These values of k:x(app) and k:x are smaller than those commonly i;l2+ at platinum.and gold electrodes.316a However. reported for Fe cathodic voltammograms that are highly irreversible (half-wave potential E1/230 mV vs. s.c.e.), yielding similarly small values of k:x(app) ("10"5 cm s-l), have recently been obtained at platinum and gold in perchlorate media from which halide impurities had been 316b rigorously excluded. The larger values of k:x(app) are therefore 313 due to the presence of halide-catalyzed, possibly inner-sphere pathways. Rate measurements at dropping mercury electrodes are not susceptible to such difficulties since the surface is continuously renewed and adsorbs most anions much more weakly than do noble metals. 3+/2+ 3+/2+ The electrochemical reactivities of Feaq and Ruaq provide an interesting comparison. At the formal potential for Ruiz/2+ in 0.4 .§,KPF6. -20 mN vs. s.c.e., the observed rate constant for electro- 3+ aq is 0.15 (10.05) cm s-l. This value is only mod- 3+ aq reduction of Fe erately (30-fold) larger than that for Ru potential, 5(32) x 10-3 cm s-l, despite the enormous cathodic over- reduction at the same potential (515 mV, corresponding to an equilibrium constant of 5 x 108) for the former reaction. Since these rate constants were obtained under the same conditions and the reactants are of very similar structure, the work terms should be essentially identical. Any reason- able driving force correction for Fez; reduction therefore must yield a value of kzx ca. 103-fold smaller than for Rug; reduction, irrespective of the work term corrections upon the individual rate constants. From h Equation 6.7, this results in a corresponding estimate of kex that is ca. lOS-fold smaller for the former reaction (vide infra). 3. Rate Constants for Electron Exchange from Homogeneous Cross-Reaction Kinetics Table 6.2 summarizes pertinent rate and equilibrium data for the 3+/2+ acid-independent pathways for cross reactions involving Feaq , Ru3+l2+. V3+l2+, Eu3+I2+ and Cr3+l2+ aq aq a a for these couples resulting from the application of Equation 6.2. The , together with the values of k2: 314 . . . . as m m on ex... 33.2.. _ m2; A e253; m 12 rm e ._~.3...m=£ a :53. . . . e m m 2 .35 e2 x... a e3; m 262:3: _ e3 .~”¢..n.a a :53. . . . . e m e e S... 2; ~23 a 1:: N2 XN e :3... Aeeemeeu s I e O m a ”a S 2 e e2; e3: eSVeSxee ~33 .Na :2”. «.mn m a a e o e = N mm a TB xn NB 3 nazNS; N a: re m .~>..m e S m1 . E 7 C _ . _ . ere... nieces. v- x moo : m x cc~ x~ H mo xN m m: x m _ new +m . . . . e m m.2332 2... {Size me_xm_ a: Sea: n eSx~_ .Nee..1ee;&eu 2;: . axe—xcze See 3.8 fixed 2.1.13. were. 2.35.: m- an e e e e. e m mm .m 1 . z . . . e m 72.; 3 e {3: a n3; ”.3 Emil _ 63 .Nee..m2e£mo 235 . _ _ a; a: 3:6 2;; maeeeeeee.eee._ ~- 3 s am no x e N 0 .~ . mm. . . . . e m e. Nb. .8 _ 2 e .131) a: XN secret; a :2; e Li; a __z:.e._.eme._ . . . . . xam n a me: re 2 e K: rm _ mg: a 2:3; m :3; _ .N :_ne=e.mmem . m 6 Tie... we... 63.1 m2 xx Semi; ~ m3; .Nse..me.._ . . . one one 2 o _ 2 3... 8e o a .N “:3 e .8 S 2 ea 2 2 eeofiueee emote e as. an r es e. :3. ea 3 ea e s. .moe cannons -mmon soon a non once ox - a menu new . . comm e e e a . e on am e a utmLTm .2; a m a m a a I X” s e +N\+n> +~\+n:¢ e~\+mom mam oemN on an own H-2V a; muemumcou oumz mouuonnou-xuox mo eomumemumw N o mmm N-e N N m-e xe Ne m N .N m ..M e_ x_ a.e ea xn ea XN ee>_. nemamzzeae e m ee .Ne .N O I a n ee— xm _ anemo— xe e_ea rm N .m>...maeeeeeeu . . m m me_ XN ANeNe_ re . Ne. rm _ wN>...Na»eeeou C O a O m eea.rN Am Names rm _ mes x. .w>...na =ze=e O I a M m.:. rm N Race-e~ xN m N .N> ..nAeeweu . . m m nee re aseNe_ xN e e N _ . .Neeeeo...m> 2; «A: 2.3.. e.— ee>.ee> N. N .N .N moacHucou N.o OHan 316 .oufiafiaauouacowfi n am“ .maaaounucmcmsauoa.a u cmna .mafivaumawnl.~.~ n >mn .mcwsamwvmamamnum u am .mawvfiuma a mag H.o coaumsum wcwma mmx .me .NHM mo mm=Hm> mafivsomwmuuoo Scum cmawmuno .maasou xovmu oscm mo mwcmsoxo mama Mom uamumcoo mumu wmuomuuoonxuozv .oH.o coaumscm an cmafimmw .uouumw muuom wafi>auv afiumucmsom .xnoa manu Eoumo .Auxmu mmmv naumafiumm mum Aalm 2 OH x HV mmaaaoo nahmnvz :fiamlsca How mmsam> .vmuoa mmwzumnuo mamas: Ham vacuummwu Scum :mxmu Nu a nu +N\+m .uamuummuou you x no m=Hm> vMuomuuoonxuozr .HHm mucoummmu cw vmawauso mm maumu xuo3 mom aowuumuuoo umumm : .cguommu mmouo How ugumaou mamas 68 muamummmu Scum .on moamummmm .mmm mucmummmu scams. .qmm N « vacuumwwu acume .nmm vacuummmu Baum” .Hmm.mo:mummmu_aoumm .vmumum mmfi3hmnuo mamas: Ham mucmwmmmu ca kuoac mmuuaom Woum :mxmu mmaam> .mmmmnuamuma aw ao>ww camcouum UNQON can oomu um coauummu mmouu vmumfia How uawumaou uuwu cmuammmmk Ammm moamuommuv H.o u : ..m.u.m .m> >5 owe: a +~\+Mahanvuu Mom mum .Aqom muamumwmuv Scum,NV @wm wucmumwouv H6 u: ..m.u.m .9? >8 mac n +N\+M¢33mo Mom mun .qw wuamumwmu 80“.; .vmuo: mmasumnuo mamas: mm wocmumwmu Baum :mxmu mmamaoo umzuo now mmaam> "H.o wanna cH cm>fiw mmansou osvw now a mo mm=Hm> .muamamusmmma mowumafix may now vaOHaam AHI~.ou:v mnuwamuum case“ no mwamu man m :a ham>wuumamwu cowumvfixo can aowuuavou mcaowumcaa mmaaaou Movmu msu mo mHmHuawuom awaken osu mum xom cum to“ . No vmu NH H mm mums: A mm: mmvasm\mv I MdH scum vmumH90Hmo .aoauummu amouo pow udmumcou aawunaawscm Hmauomu N.o mNama ou mmuoz 317 rate constants for the cross reactions, k?2’ and for self exchange of the various coreactants, 1:32, listed in Table II are taken from literature data; they are corrected for Debye-Huckel work terms as described in reference 311 (see the footnotes to Table 6.2 and reference 311 for the data sources). The measured values of kEZ’ k?2(app), are also listed, with the ionic strength at which they were determined in parentheses. The equilibrium constants K12 given in Table 6.2 were obtained from measurements of formal potentials for the appropriate redox couples at ionic strengths comparable (0.1 0.2). - . 2+ 2+ 3+ 3+ . 3+ 1nvolv1ng Vaq, Ru(NE3)6 , Co(phen)3 Ru(NE3)5py , and Ru(Nfl3)sisn as coreactants, vary by a factor of almost 200 (Table 6.2). However, the l relatively low value (0.4‘gf s-l) obtained using Co(phen);+ is also characteristic of cross reactions involving this oxidant with Viz, Eu2+. and Cr2+ (Table 6.2). Also, the estimate of kh obtained from aq aq ex 1 the Ru3+ - Vi; reaction (z‘gf s-1) is likely to be too small since the 3Q free energy barrier for Vi; oxidation appears to respond to changes in driving force to a noticeably smaller extent than predicted from Equation 6.2. The remaining cross reactions yield a reasonably consis- tent estimate of kgx for Ruiz/2+ of ca. 50(120)§._-1 s-l. (A somewhat larger estimate of kzx, ca. 200 gfl s-1 319 . was arrived at previously from but involving extrapolation of values of k: tained from cross reactions having highly varying driving forces.304) related arguments, x ob- The "observed" value of kzx for Viz/2+. 3 x 10-2 firl s-l, (i.e. that obtained directly from the observed self-exchange kinetics) is close to the estimates obtained from cross reactions with Co(en)§+ and 3+/2+ D Ru(NE )3+. In contrast, the "observed“ value of kh for Fe 3 6 ex aq . is considerably (104- to 105 fold) larger than those derived 15 ”-1 s-1 from cross reactions having suitably small driving forces (f>0.2) (Table 6.2). Equation 6.1 can be rewritten as h 12 h h log k 0.5(log k11 + log kzz) + 0.5 log K 2 12 + (log K12) / h h 2 4 log(k11k22/Ah) (6.8) 320 Figure 6.1. Plot of (2 log kh - log'khz ) vs. [log K + (log K12)2/ 12 4 log (kh / )] for homogeneous cross reactions involving Fe3+l2+ 11k 22 2h aq ’ 12 calculated as Fe:q oxidations. k2x for low spin M(III)/(II) polypyridines taken as l x 109 £71 sec.l (see text). kh refers to Fe:+/2+self exchange; 11 kgz to self exchange for coreactant couples. Zh assumed to equal 1 x 1012 gfl sec-l; value of kgl required for plot obtained by iteration: kh - 11 10'.3 M71 sec-1. Data sources given in Table 6.2 or reference 311 unless otherwise stated. Closed points refer to cross reactions; Open point to 3+/2+ 3+ 3+, "observed" Fea self exchange. Key to oxidants: l. Ruaq; 2. Euaq, 3. or3+; a. 3+82+; 5. v3+; 6. Ru (NH 3)2 ; 7. Ru(en)2+ ; 8. Ru(NH3)spy3+; aq 3+ 9. Ru(bpy)3+ 10. Ru(NH3)Snic3+, reference 320; ll. Ru(NH3éisn ; 3 ; 12. Ru(NH3)4bpy3+, reference 310; 13. Os(bpy)g+, reference 334; 14. Fe(bpy)§+, reference 335; 15. Fe(phen)2+, reference 336; 16. Co(phen)2+, reference 337; [17-19 from reference 330]; 17. Ru(terpy)3+; 18. Ru(phen)§+; 19. Ru(bpy)2(py):+; [20-23 from refer- ence 338] 20. Os[5,5'-(C33)2bpy]§+; 21. Os(phen)2+; 22. Os(5-Cl-phen)2+, 23. Ru[5,5'-(CHB)2bpy]§+; [24-31 from reference 339]; . 24. Ru[3, 4, 7 ,8-(CH3)4phen]2 ; 25. Ru[3,5,6,8-(CH3)4phen]2+, 26. Ru[4, 7-(CH 3)2phen]: +; 27. Ru[4,4'-(CH3)2bpy]§+ 28. Ru[5,6-(cu3)2phen]2 ; 29. Ru(S-cu3phen)2+; 30. Ru(s-cénsphen)2+, 31. Ru(S-Cl-phen)3+ ; 32. Fe:+. 321 ~30- -20. .c a A: O s -10 - gxg :2 5;, IS 20 - o - 22" ’2' N I7,l§39gfi‘ :22 £32 9 3° 23 {a IO+ l l I 1 I IO 0 -lo -20 -3o 2 h h 2 log Klz-t-[Oog K'Z) /4log(k"k22/Zh)] Figure 6.1 322 h Therefore a plot of (2 108 klz 12 h h log(k11k22[A§)] should yield a straight line of slope l.0 with an intercept equal to log kgl’ Figure 6.1 shows such a plot for 32 i;/2+, formally expressed as Pei; oxidations. h 2 - log kzz) vs. [log K + (log K12) [4 reactions involving Fe (The data sources are summarized in the footnotes; Ah is assumed to 12 “fl -1 equal 1 x 10 s . The value of kgl in the last term in Equation 6.8 was obtained by iteration; for most reactions choosing any reason- 4 l -l able value of Rh in the range ca. 10- to 10 gf s led to essen- ll tially identical results.) The straight line of slope 1.0 in Figure 6.1 provides the best fit to the solid points which refer to the cross reactions. The intercept, which equals log kg: for Fe:;/2* ponds to k2: - 7 x 10'4‘gfl s-l. The open point, which refers to the , corres- "observed" value of k2: (ls‘gfl s-l), is clearly at variance with the other points; yielding a discrepancy of over 104 -fold in k:x° Figure 6.2 shows a similar plot for cross reactions involving Viz/2+. Although the data points are less numerous, the open point for Viz/2+ 1 self exchange (k:x - 3 x 10-2.§f s-l) is consistent with the remaining entries. Admittedly, the slope (0.9) of the best fit line in Figure 6.2 differs somewhat from unity; possible causes are discussed elsewhere.249 3+/2+ Such a behavioral difference between Vi+l2+ and Feaq 310 has been noted previously. 3+/2+ Fea The striking discrepancies with Equation 6.2 for were ascribed in part to especially facile reaction pathways for self exchange of the cross-reaction partners that contain pyridine- type ligands arising from interpenetration of the pyridine rings. Since such interactions will be absent for the FeiZIZ+ cross reactions, 323 h k22 h 'z-log 2log k no . 5 <3 -5 -IO log Kl2+[(log Kl2)2/ «og(lfi'|k;2/z§ )] ++ Figure 6.2 As for Figure 6.1 but for cross reactions involving v2 /2 , expressed as Vi: oxidations; kql refers to Viz/2+ self exchange. Data sources From Table 6.2 or reference 311 unless otherwise stated. Key to . 4+. 3+. 3+. oxidants. 1. an, 2. Rnaq’ Ru(Nl-l3)6 , 4. Ru(NH3)5py 6. Co(bpy)2+; 7. Ru(NH3)Sisn3+, reference 321; 8. Co(phen)2+; 9. V 3+; 5. Co(en)2+; 3+ aq' 324 the resulting estimates of *2: for Fei+l2+ (ca. 10"4 to 10-.3 §_s-1) were considered to be falsely small, the observed self-exchange rate 310 constant being presumed to reflect a “normal" outer-sphere pathway. A difficulty in comparing data for cross reactions involving FeiZIZ+ with the other aquo couples is that the formal potential for Fei+l2+ is substantially more positive of those for the remaining aquo couples. Therefore cross reactions having suitably small driving forces inevitably involve different coreactants with the likelihood that systematic errors in the applicability of Ehuations 6.1 could occur. These errors could vitiate its use for obtaining even relative 3+/2+ values of the self-exchange kinetics of Feaq with respect to the other couples. However. the comparison of the kinetics of the Os(bpy)g+-Fei; and Viz-Cflbpy);+ cross reactions provides a way of circumventing this problem. Both these reactions have suitably small driving forces [equilibrium constants of 80 and 1.2, respectively (Table 6.2)] and the coreacting redox couples, Os(bpy)2+/2+ and Cr(bpy)§+/2+. have not only the same ligand composition but are also likely to have similarly small barriers to electron exchange since they both involve electron acceptance into a delocalized t28 orbital.310 3+/2+ 3 self exchange, k22 and k44, are equal and noting that the Assuming therefore that the rate constants for Os(bpy) )3+/2+ 3 and Cr(bpy f terms are essentially unity, the ratio of the rate constants for 3+/2+ h h Fea ex,Fe/kex,V’ can be found from the and Viz/2+ self exchange, k kinetics and thermodynamics data for the corresponding cross reaction using (cf. Equation 6.1]: 325 you uonu ou uooamou new: xmx mo moans“ mum numosumvas mononucouma a“ mosao> .moauosax mwcmnoxolmamm aoum :zauoouwv: vocamuno mumxomun 3% N Wmom was wm> now mosaw> .Auxmu momv H.o cowumsvm wcams A~.o_manmav sump coauomoulmmouo Bonn vocaawm 0v .owsmnwmm mama msomaowoao: you usmumaoo sump vouuouuoolxuoz mwmuo>< a .H.o manna aoum moomwuousw msoosvml>usouoa um swamnoxo Hmoaaozoouuomao you unnumsoo mush cmuoouuoolxuo3o an an an new Aoauwm x we Am-mm x “MN Amuwm x we oH x N cmuo s- a- o- o- +N\+m com Amuowaxxmwav Ac-mm H We OH x m swam «I «I ml +N\+m Aquoa x m.mV finned x m-v Am ov .AmnonN-v cm x a . x : a owe as suoa ov Hmfl_ muoa a sled x av +N\+mmm A soaxm.~v As oaxov.x uoaxav as N N N x mma Naca o --oa Me «nos n m-oa a +~\+m> . adv Aav Aav oma as x m.~ mm on oH x swam m N- N +~\+m HIaoa Hummv 5 Hum Hmm. Hum mm. Hum Mm. Hum Eu maasoo momma on o so am . no no as m owe ex a v as o A“ we as a as 8 we .maoauowvoum Hmofiuouomca sues comaumaaoo was .00mm on mwsmnuxm souuomHm now museumsoo oumm wo >umaasm .m.o magma 326 .mm.oomwuowwu scum nwxmu mosHm> .m.o: : nuwsouum oHsOH um wosHa Iumuov .oHasoo xovmu now haouuao GOHuomomm . + am How many on uomammu nuH3 Aonov xmx mo moHumu mum numoauovaa mommsumwuma aw moch> .Auxmu momv mmHMMmu moamou .mm.o was m.o m:0Huo=vm Bonn uoaHsuno mum mou< « can mwo< skews .m ou< + mHo< I xou< use .HI§ HI: NHOH x m. m I < .H u u umnu awsHasmmm N. o coHuwsvm Bonn vmustono xmx moa mosHa>m .+~\+m:m now umsu ou uomamou SUHS Ac. 0 GOHumsamvx ms mo mOHumu mum snowshoes: mmmwnuamuma sH mous> .Auxmu ommv H Hoam .Hmox o. m u M was HIoom so nOH x H n o<. Hloom HIz NHOH N m. m I dd uosu wsHasmms e. o soHumswm msHmsx Mx mo mous> waHvsonmouuoo aouw xmx mo mmumEHummo .+~\+m:m 327 h h _ , kex’Fe/kex’vhku/k”) (klz)zqu34/’-unxu (6.9) 3 1 -l a" s 3 Inserting the experimental values k12 . 3 x 10 , k34 H-1 s-1 ._ , K12 - 80, K34 - 1.2 (Table 6.2), into Equation 6.9 yields h h h l - ’ -2 - -1 . ex,Fe ex,V 0'1," Taking kex’v t° be 5 ‘ 10 E 3 (9-4—9 m} h 3 . _ - -l -l 322 . h yields the result kex,Fe 7 x 10 g s . This value of kex for +/2+ q self-exchange kinetics. . l x 10 /k Fe: is again over 103 -fold smaller than that obtained from the 4. Cogparisgn Between Electrochemical and Homogeneous Exchaggg We Strong evidence supporting the validity of such smaller estimates 0f kg‘ Fe is obtained from the rate constants for electrochemical 9 exchange, kgx. Table 6.3 contains a summary of the "best fit" values of kg: and kg: for each redox couple, along with estimates of kh kh ex’ ex (Equation 6.6), obtained from the corresponding values of kg: using Equation 6.6. The values given in parentheses are ratios of kg: and k2x (Equation 6.6) with respect to those for Ruiz/2+. (kh lkh ex ex,Ru ). The frequency factors Ab and Ae required in Equation 6.6 were estimated 10,18,186,323 from an "encounter preequilibrium" model using the . 2 3 186 . expressions Ah - 41rth 6rh vn/lo and Ae - Grevn, where N 18 Avogadrofls number, rh is the average distance between the homogeneous redox centers in the transition state, Grh is the approximate range of encounter distances (“reaction zone thickness") within which electron transfer occurs, are is the corresponding reaction zone thickness close to the electrode surface, and “n is the effective nuclear activation 328 Figure 6.3. Comparison of rate constants for electrochemical exchange at mercury-aqueous interface, kzx (cm sec-1), with corresponding rate 1sec-1), taken from constants for homogeneous self exchange, RZx (yf Table 6.3. Closed points refer to values of REX obtained from homo- geneous cross reactions; open points to those obtained from measured self-exchange kinetics. The straight line is the relationship between log ke and log kh expected from Equation 6.6. ex ex 329 frequency.23 18,10’186 Inserting the anticipated values rh I 78, 5th I 5ré _ 1 x 1013 -1 10,23 vn s for the present aquo couples into these expressions yields Ah I 3.5 x 1012,§f1 s"1 and Ae I l x 105 cm 8-1. The value of C in Equation 6.6 was estimated to be 3.0 kcal. mol-1 by inserting the values Rh I 7 2, Re I 13 2324 into Equation 6.5. (Note that although there is some uncertainty in the appropriate absolute values of both Rh and Re. this partially cancels in Equation 6.5). The absolute as well as relative values of k:x (Equation 6.6) are seen to be uniformly in good agreement with those values of 1‘21: (Equation 6.6) obtained from homogeneous cross reactions. This is also illustrated in Figure 6.3 as a plot of log k:x against log kzx. The straight line represents the correlation predicted from Equation 6.6. The solid points refer to values of kzx obtained from cross reactions, and the open circles represent the values of kg: for Viz/2+ and Fei+l2+ obtained from the self-exchange kinetics. Although all the other entries are consistent with this correlation, that obtained from the Fe3+l2+ self-exchange kinetics is again about 104 -fold larger than expected. 5. Qorrelation‘g§_ln§rinsic Barriers gig; Regctant Structure As noted above, it is instructive to compare the variations in the experimental values of kg: and kg: with the structural properties of the redox couples. To a certain extent, the observed reactivity sequence Eu“,2+ > viz/2+ > Fe3+l2+ > Eu3+l2+ ) Cr3+’2+ is consistent aq aq aq 8Q with structural expectations; the three most reactive couples all 330 involve the acceptance of the transferring electron into a t28 orbital for which the required distortions of the metal-ligand geometry, and hence AGES. are anticipated to be relatively small.309 It is stressed that the calculations to follow are only approximate. with more detailed calculations to appear in Chapter VII. The primary purpose here is to discover the relative reactivity se- quence that is predicted by electronrtransfer theory on the basis of structural information. The absolute values of calculated rate con- stants may well be in error by one to three orders of magnitude due to uncertainties in the values of critical parameters (e.g. metal-ligand 'stretching force constants) as well as to simplifications introduced in the calculations. Nevertheless, such calculational errors should be closely similar for each redox couple, so that a useful prediction concerning the sequence of reactivity can still be obtained. The calculation 0f AG; and hence ktelx or ke from electron- ex transfer theory requires quantitative information on the changes in the metal-ligand bond distances, Aa. accompanying electron transfer.13’23 Although sufficiently reliable determinations of As are sparse, recent X-ray diffraction measurements have established values for Ru(0n2)2+/2+ 3+IZ+ °f 0'09 211 and 0'14 2.325’326 respectively. An 2)6 effective value of As for Cr(OHz)g+/2+ equal to 0.202 has also been 23 and Fe(OH determined from solution EXAFS measurements.10 The relation as; - nf2f3 mnznuzws) (6.10) 331 where n is the nwmber of metal-ligand bonds involved (six) and f2 and f3 are the metal-ligand force constants in the divalent and trivalent oxidation states, assuming that the corresponding metal-oxygen stretch- ing frequencies are 390 cIII-1 and 490 cmfl, respectively, yields values of AG? of 14.6, 35 and 72 kJ mol"1 for Ru3+l2+ Fe3+l2+, and Cr3+/2+, is aq aq aq respectively. Errors in these values may arise both from anharmonicity of the potential-energy surfaces as well as from uncertainties in the . . * appropriate force constants. The effective values Of‘AGia are slightly smaller as a result of nuclear tunneling23; the nuclear tunneling fac- 3+/2+, Fe3+/2+, and Cr3+/2+ aq 3Q 3Q yield effective values of A028 equal to 13.6, 33.5 and 67.4 tors, Tn, of 1.5. 2 and 6 for Ru 10,23 , respect- ively, -l . . . a * kJ mol , respectively. The outer-shell contribution to AGex’ A603, can be estimated from Equation 6.4a again using the values a I 3.5 X, a - Rh I 7 8 yields AGO8 I 27 kJ mol 1. Inserting the resulting estimates of Aczx into Equation 6.2 along with the above estimate of Ab, 3.5 x lOlngrl s-l, and assuming that K I 1 yields the calculated values of kh to that for Ru k:x(calc), listed in Table 6.3. Ratios of k:x(ca1c) with respect 3*’2*.8 kcal mol-1), 358 * alongside the corresponding experimental values of ACh 12. It is seen 9 I that the estimated values of AG diverge from the straight line h,12 predicted from the harmonic oscillator model to a similar. albeit slightly smaller. extent than the experimental values. Admittedly. there is no particular justification for assuming that the reduction half reactions obey the harmonic oscillator model. However, it turns out that the estimates of AG;,12 are relatively insensitive to altera- tions in the shapes of the mil - AG: plots. It therefore seems reasonable that the deviations of the activation free energies for highly exoergic electrochemical and homogeneous reactions illustrated in Figures 6.5 and 6.8 may arise partly from the same source, i.e. from values of oz for the oxidation half reactions that are unexpectedly small. That is not to say that other factors are not responsible, at least in part, for these discrepancies. Nonadiabaticity, work terms, specific salvation, and other environmental effects may all play impor- tant roles depending on the reactants. For example, there is evidence 3+/2+ to suggest that the true rate constant for outer-sphere Feaq self- exchange is significantly smaller than the directly measured value (see Section VI. A). This can account for a good part of the unexpectedly slow rates of cross reactions involving this couple. It is remarked here that in the original publication we were unable to uncover the reasons for the peculiar driving-force dependence of the aquo oxidation kinetics, although a number of ideas were advanced. However. absolute calculations reported in Chapter VII indicate that the driving-force anomalies on both the anodic and cathodic sides for electrochemical reactions of Fig/2+ and Griz/2+ can 359 be attributed in large measure to differences in inner-shell force constants for the M(III)-L versus the M(II)—L state. Such differences also should account for the homogeneous reactivity patterns. Rather than indicating a failure of electronrtransfer theory, these findings point to the unsuitability of the "equal force constant" approximation which is employed in deriving Equation 6.24, as well as 6.25 and 6.28. This discussion therefore illustrates the inadequacy of the relative theory in comparison to the absolute theory for solving one aspect of the problem. However, in another regard the relative theory is the superior approach in that the reaction energetics of redox couples such as Eu3+/2+ sq can still be examined even though structural data (required in the absolute approach) are lacking. Furthermore the relative electronrtransfer theory appears to represent the only method of inter- preting kinetics data for chemically irreversible reactions for which thermodynamic data are lacking. Thus, the two approaches are comple- mentary in generating insights and understanding concerning the details of electron-transfer kinetics. 6- mm It seems clear that kinetics as well as thermodynamics data gathered for simple electrode reactions can contribute significantly towards the development of our fundamental understanding of electron transfer in condensed media. In particular, detailed studies of electrochemical kinetics with due regard for work term corrections can yield information on the shapes of free energy barriers and the isolated “weak overlap" reactivity of metal complexes. 360 C. Entropic Drivipg Fprcg Effects Upon Preegppnential chtors for Intramolecular Electron Tppnsfer: Ipplications for the Assessment,p§,§pnadiabaticity (Accepted for publication in Inorg. Chem.) Considerable effort has been directed recently towards measuring rate parameters for intramolecular redox reactions, typically involving electron transfer between a pair of transition metals linked by an 353-355 organic bridge. Such reactions offer important advantages over bimolecular processes for investigating detailed aspects of electron transfer. since the uncertainties concerning the energetics of forming the precursor complex are absent and the geometry of the transition state is relatively well defined.355 The study of intramolecular electron transfer should be parti- cularly useful for examining electronic coupling effects; i.e. the factors influencing the occurrence and. degree of 353b-d,355b nonadiabaticity. Since the rate constant, k for such et’ processes should refer to an elementary electronrtransfer step, it can be expressed as342 It“ wn Kelexp(As;C/R)exp(-AH;CIRT) (6.29s) ket I\hexp(A8:/R)exp(-AH;c/RT) (6.29b) where vn is the effective nuclear frequency factor, K e1 is the * electronic transmission coefficient, and ASFC * and AflFC are the entropic 361 and enthalpic components of the Franck-Condon barrier. Values of Kel can therefore be extracted from rate measurements as a function of I: temperature provided that vn and A8 are known. For many reactionsv 11 PC is numerically very similar to the conventional factor kT/h.9’23 Since work terms are absent for intramolecular reactions, the observation of preexponential factors that are substantially less than vn, or equi- I: valently, of negative apparent entropies of activation ASa is often taken as prima facig evidence of nonadiabaticity (i.e. meld). I: Implicit in this interpretation is the assumption that ASFC is essen- tially zero. However, contrary to common belief this is probably incorrect for most nonsymmetrical intramolecular reactions, including those featuring charge symetry. The purpose of this section is to explore the likely magnitudes of AS* for some representative intra- molecular reactions and to examine the consequences for the interpre- tation of unimolecular preexponential factors in terms of nonadia- baticy. The origin of the Franck-Condon activation entropy can be clari- 13 for the activation free 356 fied by considering Marcus’ expression I: energy, AGFC’ and its temperature derivative. Thus, * * AG I AG. t + 0.5 (1 +OL)AGO (6.30) FC in I: where Mint is the intrinsic barrier, AG0 is the thermodynamic free I: energy change, and o I AGO/82361“. The temperature derivative of Equation 6.30 is 362 * * 2 O Ach I AShun-4o ) + 0.5 (1 + 2a)AS (6.31) * . . where ASint is the intrinsic activation entropy and AS0 is the entropic driving force. For reactions having small or moderate values of Co, I 0 so that Equation 6.31 simplifies to * * o ASFC Asiat + 0.5AS (6.32) * For reactions in aqueous media, Asin 1 t is usually assumed to be close to zero (within ca. 2 J deg- mol-1) as predicted by the conventional dielectric continuum expression,203 although as shown in Section V. A there is good reason to suspect that small positive values 1 I _ - of ASint (ca. 4-12 J deg mole 1) are common.357 However, A80 can often be sufficiently large to yield a substantial contribution to * Ach' even for reactions leading to no net change in ionic charge. For example, the reduction of Fe(phen)2+ (phen I phenanthroline) by Fe(OH2):+ yields a net entropy change of -l67 J deg”1 mole-1,55 l * - - corresponding to a value of SPC of ca -67 J deg mole 1 from Equation 6.31. The values of AS0 can be related to the component entropic chan- ges (the so-called "reaction entropies" As:c)55 at the redox centers undergoing reduction and oxidation, As: and As:c ox’ respectively, 9 c,red by O O 0 AS - Asrc’red -ASrc ox (6.33) 363 Such large entropic driving forces arise because the component reaction entrapies are often large and extremely sensitive not only to the charge type of the redox couple, but also the chemical nature of the coordinated ligands, the metal redox center, and the surrounding sol- 52.82.84 vent. This is due to the large changes in the degree of specific ligand-solvent interactions, and hence the extent of local solvent polarization, brought about by electron transfer.55’82’222 I PC were calculated for thirty outer-sphere bimolecular reactions and com- In order to check the validity of Equation 6.31, values of AS pared with the experimental activation entropies, Asz‘p. The choice of reactions was limited to those for which values of 48° and AG0 are available and sufficient structural information exists to permit the calculation of AGE“. Details will be presented in chapter VII. All these reactions involve reactant pairs having charges of +3 and +2. * * 358 against AS . This plot Figure 6.9 is the resulting plot of ASm exp demonstrates that the experimental activation entropies do indeed respond to the entropic driving force roughly in the manner predicted I by Equation 6.31. although the values of Asexp are about 60 to 100 J -1 1 - I deg mole smaller than AS . (Also note that reactions involving PC high-spin Co(III)/(II) couples do not display noticeably different behavior in spite of the spin state change that occurs for these reactions.) This discrepancy could be taken as an indication of nonadiabaticity (Kel (<1), but more likely it arises chiefly from an additional component of [58sz associated with the unfavorable entropic work of forming the highly charged precursor complex from the separated cationic reactants. 364 Figure 6.9. Plot of the experimental activation entropy, ASpr’ against the activation entropy, AS§C, calculated from Equation 6.32 for bimolec- ular reactions involving 3+/2+ redox couples. Data sources quoted ref- erence 311 unless otherwise noted (aq I 032, an I ethylenediamine, bpy - 2,2'-bipyridine, phen I 1,10'Iphenanthroline, terpy I 2,2',2"-terpyridine py I pyridine). Reactions: 1. Co3+ + Fe2+; 2. Co3+ + Cr2+; aq aq aq aq 3. Fe3+ + 1162"; a; 5.3+ + v“; 5. $433+ + Fe2+; 6. v3+ + v“; a aq aq aq’ aq 8Q aq 3Q 3““ 2+. 3+ 2+. 3+ 2+. 7. Feaq + Ru(NH3)6 , 8. Fe q + Ru(NH3)5py , 9. Feaq + Ru(en)3 . 10. Ru(bpy)3+ + Fe2:§; 11. Ru(phen)3 +Ru 2:; 329 12. Fe(bpy)3+ + Fe 2+ 359 13. Fe(phen)g+ + Fe::;335 14. Ru(terpy)2 + Fe 2:; 329 W 3+ 2+.329 2+. 3+ 15. Ru(bpy)2(py)2 +Feaq, 16. Ru(NH3)6 +Vaq.17. Ru(NH3)5py *- 2+ 3+ 2+. 3+ 2+.31o 3+ Vaq’ 18. Co(en)3 + Vaq, 19. Co(bpy)3 + vaq’ 20. Co(phen)3 + 2+.310 3+ 2+.360 3+ 2+.28 vaq’ . 21. Ru(NHa)5py + Crag, 22. Ru(NH3)4bpy + Ru(NH3)4phen , 3+ 2+. 3+ 2+. 3+ 23. Ru(NH3)6 + Ru(NH3)6 , 24. Ru(en)3 + Ru(NH3)6 . 25. Co(en)3 Co(en)§+; 26. Co(bpy)3+ + Co(bpy)§+; 27. Co(phen)§+b+ Co(phen)§+, 28. Co(en)3 + Cr(b py)§+;361 29. Co(phen);+ + Ru(NH3)§+; 30. Co(phen)§+ + Ru(NH3)5py2+. deg" mole" asap, J 365 4.0% (18%;, J deg"mole" Figure 6.9 €53 s 5 O 6 20 O in '5 02: ITO 0 019 Ono 27 n O ' '3()0 l6 0 :2 o 0'4 L 1 1 -25 O 25 366 Most nonsymmetrical intramolecular reactions will also have nonzero entropic driving forces, thereby yielding nonzero values of 38:0 on the basis of Equation 6.31. This is most obviously the case for systems where electron transfer leads to a net charge decrease at the redox centers, such as the pentaamminecoba1t(III) - pentacyanoiron(II) reactions examined by Haim and coworkers.354 Here the values of Ass-c,red and Asicmx will be positive and negative, respectively (Table 6.4), yielding large positive values of A80 (Equation 6.32) and hence A8;c (Equation 6.31), due to the decrease in solvent polarization attending such charge neutralization.354C Quanti- tative values of AS0 for these systems are difficult to estimate from values of As:c for the isolated redox centers. This is because As:c for cyano Fe(III)/(II) couples are known to be extremely sensitive to the cationic environment (Table 6.4)286 so that the proximity of the cationic Co(III)/(II) redox center is expected to influence the solva- tion around Fe(III)/(II). Nevertheless, from the As:c values in Table 6.4, we estimate that As:c equals roughly 125 J deg.1 mol.1 and I45 J -1 mol.1 for the Co(NH3)5L-3+/2+ and ILFe(CN)§-/3- fragments, where L is a pyridine-type ligand, yielding A8°==170 J deg.-1 1 deg mol.1 (Equation mol.1 (Equation 6.31s). Indeed values of I . Asa of this order are obtained for (CN)5FeII-L- CoIII(NHB); reac- 354 6.32) and AS;C :80 J deg- tions. Similarly, large values of A30 are expected for the intra- molecular reduction of pentaamminecobalt(III) by the p-nitrobenzoate radical anion since both redox centers will again contribute positive components to A80, thereby explaining the large preexponential factors I (i.e. positive values of A88) observed for these reactions.362 Reaction Entropies, A30 (J deg- 367 Table 6.4 l mol-1) for Various Redox Couplesrin Aqueous Solution.8 Redox Coupleb M(Z-bpy)g+/2+ u(puen)g*’2* )3+/2+ 2 6 )3+/2+ 3 u(os M(en 3+/2+ u(nn3)6 M(NE3)5(bpy)3+/2+ M(Nfl3)5Py3+/2+ n(nu3)50s2 M(Nfl3)5NCS 2+/+ M(Nfi3)561 2+l+ c-H(Nfl3)4Cl2 u(cu):'/4' 3+/2+ 2+l+ H(¢.‘.N)4bpy-/-2 Co 92 92 250° 155 190 Metal Center Ru Fe 4 8 150 180 54 75 71 105 63 42 42 -l76f, - 96"8 I120 368 Notes to Table 6.4 a Ionic strength uI0.1; data taken from references 55 and 84, except 1 1 where noted. Uncertainties estimated to equal 2-6 J deg- mol- for directly measured values. bl! denotes metal redox centers. Ligand abbreviations: 2-bpy I 2.239bipyridine; bpy I 4,4’Ibipyridine; phen I l. lO-phenanthroline; en I ethylenediamine; py I pyridine. cEstimated from thermodynamic data given in reference 398by correcting for tem- perature dependence of reference electrode and likely ionic strength 1 effects. A similar value (ca. 1 20 J deg- mol-1) is obtained from a correlation of 83c with the self-exchange rate constant kex’ assuming that her 10-10 5.1 sec-1. dEstimated by comparison with corresponding 3+/2+ )3 value for Co(en . eJ. T. Eupp, unpublished experiments. ronic strength uIO. 8Determined in 0.2 g La(CLO4)3. 369 Substantial values of A80 are expected also for intramolecular reactions between pairs of metal redox centers having the same charge type and similar ligand environments but with different electronic conf igurations. Important examples of this type are the pentaammine- cobalt(III) - aquotetraamine- ruthenium(II) reactions bridged by nitrogen heterocyclic ligands that have been studied by Taube and 353 coworkers. Since these reactions are charge symmetric, it is usually tacitly assumed that A80 I0 and hence As;c . 0.353b.355b However. since reaction entropies for high-spin Co(III)/(II) couples are generally found to be markedly larger than for Ru(III)/(II) and 55.84 other low-spin couples (Table 6.4), large positive values of AS0 and hence 138:6 are therefore anticipated for these reactions. As an illustrative example, we consider the intramolecular reduction of Co(III) by Ru(II) in (£120) Ru(m3)4(bpy)Co(NH3)g+ (where bpy I 4,4’Ibipyridine). The reaction for the corresponding dipenta- ammine complex Ru(m3)5(bpy)0o(un3)g+ is liable to involve a net 1 mil-1 entropy change of ca 85-105 J deg- since the reaction entropies of Co(III)/(II) and Ru(III)/(II) couples having similar ligands uni- formly differ by this amount (Table 6.4), (references 55 and 82). The effect of replacing one ammonia on ruthenium by an aquo ligand can be gauged from the 30 J deg.1 mol.1 larger reaction entropy for Ru(lfil3)50n§+/2+ than for Ru(NH3)2+/2+ (Table 6.4). (To a first approximation, the reaction entropies for mixed-ligand complexes appear to arise from linear additive contributions from each ligand).84 Thus the entropic driving force for electron tranfer in II III EZORu (N33 ) 4( bpy)-Co 1 (m3)? is estimated to be ca. +65 J deg- 370 mol-1. Bearing in mind the likely value of A821"; (3-8 J deg-1 mol-1) ,357 a value of 188:0 of ca. 35 J deg.1 mol.1 is obtained from I Equation 6.31. Therefore the "measured" activation entropy, Asa’ of 10 J deg”1 mol-1 for this system is suggestive of a significantly nonadiabatic pathway (Ke1 3 0.1; Equation 6.29) rather than the adiabatic pathway that has been inferred without consideration of the 35“": Since similar values of 133* should 8180 PC III ( m3 ) 2+ react ions with entropic driving force. . II be appropriate for other nzoau (1013) 4LCo related bridging ligands L, the smaller or negative values of As: seen, 1 for instance, with L I l,2-bis(4-pyridyl)ethylene (5 J deg- mol-1) and 1 l.2-bis(4-pyridyl)ethane(-40 J deg- mol-1) infer the presence of I decidely nonadiabatic pathways. Thus from Equation 6.4. if ASFC I 35 J -1 2 4 ~3 x 10- and l x 10- for these two reactions, mol"1 then K .. e1 deg respectively. Naturally. these resulting values of Kel should be regarded as being only approximate. Since the values of A8:c for the intramolecular binuclear reactions are inferred from data for structurally similar mononuclear couples, the reliability of the resulting estimates of AS0 and hence 138:0 may be called into question. Unfortunately. values of AS:0 cannot be measured directly for these and other thermal intramolecular reactions on account of the rapid equation that follows the formation of Co(II). However, Steggarda et al. have shown that the values of Asgc for mononuclear Ru(III)/(II) couples containing pyrazine ligands are essentially the same as in binuclear complexes where the pyrazine ligand is also bound to another ruthenium 363 redox center. This result therefore provides strong support to the 371 present method for estimating ASo values for binuclear complexes. Providing that the present estimates of A80 are accurate to within ca. 1 I1 20 J deg- mol . which seem reasonable, then the corresponding estimates of Kel are reliable to within ca. 4 fold. Such entropic driving force effects also provide a rationalization 1 mol"1 for 3+ 353d of the substantially more positive value of As; 44 J deg- electron transfer in (SO3)Run(NH3)4(pyrazine) CoIn(HH3) S 9 . . . . II III 5+ especially in comparison with that for HZORu (8H3) 4(bpy)Co (m3) , 11 J deg.1 mol-1. This increase is difficult to explain on the basis of electronic coupling effect3,353d but can easily be understood in terms of the influence of nonbridging ligand composition on A80 and hence AS;C.55'84 The presence of anionic ligands generally yields substantially smaller values of Asgc. The influence of substituting an aquo by a sulfite ligand on the reaction entropy of Hu(III)/(II) can be gauged roughly from the decrease in Asgc, 65 J deg.1 mol.1 between Ru(NH3)SOHg+/2+ and c-Hu(NH3)4Clz+/°, (Table 6.1).55 corresponding to 1 an increase in AS“r of ca. 35 J deg- mol.-1 (Equation 6.3l) since the PC ruthenium undergoes oxidation. Similar considerations can also be applied to a number of other intramolecular reactions, such as those involving nonsymetrical bridging ligands.353’365 Weaver and co-workers have also recently analyzed activation parameter data for a number of intramolecular redox reactions at metal surfaces (i.e. electrochemical processes involving adsorbed reactants) in a similar manner. Entropic driving force effects are generally important for these processes since only one redox center is involved, so that ASo equals asgc. Although the 372 estimation of As;c for intramolecular systems should be approached with caution since the factors influencing the thermodynamic entropy changes are incompletely understood, it is clear that careful consideration of the ligand composition and the chemical as well as electrostatic nature of the redox centers is required in order to evaluate the contribution of the Franck-Condom entropy to be measured preexponential factors. 373 CHAPTER VII COHEARISONS BETHEEB EXPERIMENTAL KINETICS BARAMETERS AND THE ABSOLUTE PEEDICTIONS OF ELECTRON-TRANSFER THEORY A. Introduction The kinetics of inorganic electron transfer reactions have been 366 widely investigated over the past thirty years. Experimental work in this area has been spurred by the ongoing development of detailed and sophisticated theoretical treatments of these reactions. The theore- tical descriptions of homogeneous outer-sphere reactions of transition- metal complexes in particular have reached a high level of refine- .56 .57 ment It is probably fair to say that theory has remained one step ahead of experiment since the original work of Marcus.367 Thus, in the absence of crucial structural and spectroscopic information the predic- tions of electron transfer theories in an absolute sense have in large measure remained untested. Researchers have had to be content with relative evaluations of theory. such as tests of the so-called cross relations, relative rate comparisons, etc. Very recently this state of affairs has changed. Sutin and coworkers have determined from EXAFS measurements the magnitude of metal-ligand bond distance alterations accompanying changes of oxidation state for several transition metal complexes.10 These data together 11 . 12 .325 .326 ,368-371 with previous structural measurements have been 374 used by Sutin to estimate inner-shell Franck-Condon barriers and abso- lute electronItransfer rate constants for a number of homogeneous self- exchange reactions.10 Broad agreement is found between theory and experiment. It would be interesting and useful to extend this study to electrochemical reactions. Indeed this is possible by using the newly published EXAFS data. Also it may be enlightening to examine homo- geneous crosa reactions. In comparison to directly measured self- exchange rate constants, the number of cross-reaction rate data avail- able for comparison with theoretical predictions is quite large. Thus a comprehensive evaluation of contemporary theories should be possible. Besides enabling a considerable expansion of the data set, a virtue of cross reactions is that they enable important supplemental features of the theoretical treatments to be tested, most notably the predictions concerning the response of rate constants to changes in thermodynamic driving force. Comparisons between experimental and theoretical activation parameters provide further tests of theories. In this chapter, kinetics parameters for approximately ten electrochemical and fifty homogeneous electron-transfer reactions are calculated using current theories and are compared with the corresponding experimental parameters. It is hoped that such comparisons, by uncovering the areas of agreement as well as disagreement, will indicate which aspects of electronItransfer processes are well understood while identifying problems which merit further investigation. 375 3. Calculation 9; Kinetics Pargmeters In so far as this work overlaps with that of Brunschwig and 10 Sutin, their semi-classical approach is largely retained. (Neverthe- less, in instances where commonly employed approximations produce signi- ficant errors, more rigorous calculations are made). More sophisticated (and difficult) quantum mechanical analyses are certainly possible, as 63 shown for example by Marcus and Siders. However, it has been demon- strated convincingly that the semi-classical and quantum approaches yield virtually identical results when the same assumptions are used in each calculation.23 The basic elements of electron-transfer theory have been outlined in Chapter II. According to the semi-classical treatment the overall rate constant can be written «:23 o, k - KAvnrnKelexp(-AG*/RT) (1.24) or o * k I KAvnK e1exp[--AG(T) IRT] (7.1) where K: is the precursor formation constant, Vn is the nuclear activation frequency Kel is the transmission coefficient for electron tunneling, n is the nuclear tunneling correction, 136* is the activation free energy representing the height of the classical Franck-Condom barrier. and AG*(T) is the quantum mechanical equivalent of this barrier. An additional factor which should appear in the rate formula- 376 tions is the electrostatic work of assembling charged reactants. However, work terms in both the electrochemical and homogeneous reac- tions are taken into account instead by correcting the experimental rate parameters for such work. This enables experimental results obtained under dissimilar conditions to be intercompared in a straightforward manner 0 l. Pre-expgnential Terms The precursor formation constant for electrochemical reactions is equivalent to the effective thickness of a heterogeneous reaction zone 9 (Section IV. A). A value of 6 x 10’ cm is deduced from considerations relating to the probable distance dependence of the tunneling coeffi- 15 cient Kel’ (This is discussed further in Sections IV A and C). Vhlues of I: for homogeneous reactions were estimated from:9 o 2 EA 41rN(a1 + a2) 6r (7.2) where N is Avogadrofis number. a1 and a2 are the radii of the reactants (Table 7.1), and 6r is the homogeneous reaction zone thickness, again taken as 6 x 10.9cm.372 The activation frequency vn was calculated from Equation 7.3:9 2 I I I I O I A . AGin,l+ in,2 Gin,2 A in,1+ G vn (v AG in,2 I A out out+vin,l )/(AGout+ G ) (7.3) I . . where v out and Acout are the frequency and activation free energy associated with solvent (outer—shell) repolarization, v. and ‘V. in,l in,2 377 are the inner-shell bond vibration frequencies for each of two I I . . reactants, and AG- and AG. are the corresponding contributions to 1n,l in,2 the overall activation free energy. (Only one inner-shell term appears in the calculation of v n for electrochemical reactions). A value of 9 x was used for v .9 The values of vin for the most part are out not known with certainty. since the necessary spectroscopic data concerning symetrical metal-ligand bond vibrations are lacking. The estimates listed in Table 7.1 were obtained by extrapolation from measurements on closely related systems as summarized in the footnotes. The uncertainties in v in values lead directly to only minor uncertainties in calculated rate constants. For almost all of the reactions vn values in the range from 9 x 1012 s.1 to 1.3 x 1013 a”1 were calculated. (See Table 7.2). Although some controversy exists concerning the correctness of Equation 7.3, alternative formulations17 were found to yield similar values of the frequency factor. Since values of the transmission coefficient can be calculated 14,15 satisfactorily only by means of 31; initio techniques, this parameter was simply assumed to be unity at the closest separation distance of the reactants. While this assumption represents a possible source of significant error. some justification is provided by Newton’s “"15 2+,” and Ru(NH3)g+/2+ self-exchanges as well as our own experimental work (Sections IV. C and V. C, and calculations for the Fe(H20) reference 373). These studies indicate that Kel values for homogeneous as well as electrochemical electron-transfer reactions probably lie between 0.2 and l. N moc ace ace Nee nNc use mNo OI: Ill: .33 moomvon> axon N :scc N III III nzmn a ence N III III ssz a mace N manna gnome :zmo N anon N neooc sno0n no.~ on.N nooc macaw on N do n uuomn uumnm cc.~ Nm.N women Ev.0”.an hm.~ mn.n mv.uoaNc annmm Nq.a 0N.N mmNNn muodn No.d mN.N agar "nan NIao also ANIaov (NWIEUV oczvv ocavv N> n9 cmx~m omxnm m a mi ”Ml : .Nc.c+a~.o INo.qwon.o xo xo 686.56 a X Io o x: unco.qwqa.o :~o.qw~c.c >no.QH3o.o =Hs~.o s.ano.Qflo~.o ono.qu~.o uuo.GHWm.a ouo.qwoo.o wa cud w.~ 0.5 o.~ m.N Q.N m.m m.h o.n N.¢ N.e m .modmsoo xoowm mam mmmuoammmm dmmauomuum mes oaseczvoamozh : a o No men +~\+nx 6; v o m an Na cs +N\+na avoo m a a mod enccl +N\+NA n vuu 68.73 nan aam.ou;v n n a Na ohm +N\+naeo: Von n Ia m oNN +N\+NA ovum m - may a me Noaca +N\+nA any a n I a no ANION +N\+NAco: you m an e xcqoa +N\+nA Ava: n «ma one +Nx+mAeovoo m I no a an em +Nx+nA V a o m I m an 03 +N\+nA :5 a I e N ocmN sowed +N\+on avoo I m N noN coo +N\+aao xvmu e N and can +N\+nAo :vom I o N and mac +Nx+nAo =v> o N a and ma +Nx+nAc my a HIdoa ~.s.m.u «fiasco xovom mom .5 .m> >lv NI um an A owa m .~.N edema 379 .mmamsou mmfisammxmn wcfivmommmuuoo How mmmam> Hence cu vmammm< .3: .Nmm mommmmwwm .wm .mm\NoN> wmwammmm an wwmumm mafiammxm: mafimaommmmuou aomw mmumawumm .ww mv\NoN> mafiammmm An N) Bonn mommafiumm .oo .omm oucmummmm .mm .mumm onus mundane one wmmumm +N\+mammzvz ma museum mugs zw0Hmmm ma HI So on + Aamoauuoaahmmv>naamofiuuoaahmv> wcHammmm mm mommawumm .ou .mo\NoN> 80mm mums um . .mum m as u no saw 6 a we saw» e+~\+m 1.2+on ~50 n 2 m\N o 9 HIau OH on on momma .mm . menopause mom .umucmo House mama wmfi>mn mmxmamaoo seesaw mam omum mom mm N menu wcfiuoa he +MAmmzvmm mom mommmumuw wmfinoummum ommwNNIHmuma Hmoqmuoaaxm 80mm mommaaumm .u .+N\+mA:mnmvoo Aamnmvmm mom mmam> m< Hence a» mmsmmm< .x .nmn mocmumwmm .3 oAmmzvz :a museum huNUNmoaumm nuw3 zwoammm an mesam> new mmam> v< Hence cu vmammm< .h .+N\+m .wom mucoummom .> .on one on mmucmuomom .: .Aoa mocmmmmouv mmmcono nummma econ HmHHOummvm mam Hmwxm one mo mwmmm>m no mandamummu mmam> .u .ON oommummmm .m .uxmu ma vocaaumo mm mommafiumm .m .HN mommuommm .v .+N\+mA%mnvmm Homo om< Hence om mosmmm< .m .muowmwm summoned owdoa mamxwa mam mmouuuoao wommumwmu mo mocmmmmmmm musumuomaeu mom wmfiuomumoo he .mmm mommmmmmu cw nm>fiw some afiammmvoaumnu 60mm moumaaumm .o .mmN mocmmowmm .: .Hmm moammmwmm .8 .«mm was man moummmommm .>mo.o: an Homage +wamAanvmm mam +N\+m mmsHm> mm3mafina .Hmuauaemfl mum mN\+MMWm:mV=m was +N\+mA>mnvmm mow momam> mm om N: 2H mH .H .HHm moammmwmm .x .Nmm oommmmwom .m .AM>+N 9V\N> N> "om mnemuooum ooumamoamo .nuuouum vnmwaalamuma Hmofiuumaamm mom xummmvmum vmommom we omNm>N omumaaumm .H .mouon mm ummuxm ma.n moaummum mafia: kn mommamuamo .mumuw A:ammuv:m mom fiHHvz mu ummumcou momom magnoumuum ommwfianamuma Hmofiuumaamm mo mmam> mommaaumw .n .mouoc mm ummuxm .MH.N mowummvm mean: he omumasoamu .oumum AHHHVZ ma ummummoo mouow wmfinoummum mammNHINMuma Hmowuumaahm mo o=Hm> mmumafiuwm .w .mumum AmHvz ma zommsumum wowsououum vmmeHIHMuoa Homauuoaamm mo o=Hm> mommafiumm .m .mumum aHHHvz aw hoammvmuu wagsuummum vmmwfialamuma Hmuuuumaazm mo unam> mommawumm .m .mumum moaummfixo mo owmmnu wmfixmmmaooum summed once mmmmNHIHmuma ma ommmmu .m .ON menopause Eoum mum mommaflumm .msfimmu ummuommm .u . Hoe HI HI Ion omomuumam on umm>aoo ou >8 mQN mvmvmmomuooam Hmaoamo menopaumm .m> mammoo women no Hmfiummuom Nmauom .m mom h «IN mum moaucamuumomn .mmuoa mafiamonuo woman: mm oommuowom aoum mum mama .Ammz .m> mamfiucmu 380 2. FranckICondon Barriers Three factors contribute to the classical Franck-Condon barrier, namely, the solvent reorganization energy vout’ the metal-ligand bond (inner-shell) reorganization energy vin’ and the free energy driving force AG°.13 The reorganization energies correspond to the energy required to displace the nuclear (solvent and bond) coordinates of the reactants so that they match those of the products (Figure 7.1) under conditions where AGo I 0. For a cross reaction such as A + B + A + B (7.4) the total intrinsic reorganization energies f and r for forward and reverse reactions are given by Harcusfl additivity rules as13 A 3+ B 2+ 1 :- 1 ’ A ! A f ( in + in + out) (7'5) and A2+ 33+ )\ II A.’ A,’ 1 r ( in + in + out) (7'6) where the superscripts on kin designate the particular reactant being reorganized. Equations 7.5 and 7.6 are sometimes reformulated using a pair of kin values derived from reduced force constants for metal-ligand vibrations rather than the four kin values obtained using the individual force constants for inner-shell modes. This approximation was avoided since it yields substantial errors in AG* for certain reactions, e.g. 10 11 12 13 14 15 16 17 18 19 20 21 22 Table 7.2. 381 Experimental and Theoretical Rate Constants (M-ls-l) for Homogeneous ElectronITransfer Reactions. oxidant 3+ V0120)6 3+ Fe(H20)6 ¥ Co(HZO)6 Co(HZO) Co(HZO) Co(HZO) Fe(H 0) 2 Fe(H20) Fe(H20) Ru(HZO) Fe(H20) Fe(H20) ¥°$°$°¥°¥°¥°¥°¥°$°$ Ru(HZO)6 2;: Ru(NH3)6 Co(en)?+ 3+ Ru(NH3)6 Co(en)? 3+ Co(HZO)6 Ru(bpy) 3" Ru(phen):+ Ru(terpy)§+ Fe(phen)§+ reductant «1120) 2"” Fe(H20) (2: Co (1120) 2" @0120):+ @0120):+ V(H20)§+ Ru(HZO) :1" Cr(HZO)§+ «1120)? «1120)? Ru(NH3) 3" Ru(en) :2: 2+ Ru(NH3)6 2+ V(H20)6 Fe(phen)§+ 2+ Fe(HzO)6 2+ Fe(H20)6 2+ Fe(H20)6 2+ Fe(H20)6 1.5x10' 4 8 50 1.3x10“ 9x10S 2.3x103 2.32.103 1 . 23x104 2.81.102 3.51.105 1.4X105 1.46.10“ 2x102 3x10"4 1.52.103 4.6x10’ 1.4x104 6.4x105 8x105 7 .2x105 3.7x104 (2) (0.55) (3) (l) (3) (3) (1) (1) (1) (1) (0.1) (0.1) (1) (0.2) (1.0) (0.1) (0.1) (3) (l) (l) (1) (0.5) corr 8x10- 52 33. 4.22.102 5.3x104 3. 7x106 1.9x104 1.91.104 1.5x105 2.3x103 1.5x107 3.6..106 1.11.105 5.61.103 1.6x1o' 6.5x104 1.2x10’ 3.8x104 1. 7x106 2.2x106 2x106 7.4x104 k (T) Reference 1.7 9.7 2 . 2x10" 21.105 5.2x107 3.3x109 1.1x106 4.4.405 5.61.107 3.2x10S 8.2x108 1 . 1x109 5.5x106 2.11.104 1 .4x10’ 4.8x105 5 . 2x10- 3.3x109 10 1.3x10 10 1.3x10 10 1.3x10 4x108 267 266 399 400 401 401 304 402 379 304 268 268 304 403,404 337 405 337 406 329 329 329 335 382 Table 7.2 (continued) 2§$g§2£_ .EEEEEEEEE. .3L. .JL. .EEEEE k (T) Reference Fe(bpy)§+ Fe(H20)§+ 2.7x104 (0.5) 9.6x104 1.3x109 335 Os(bpy)§+ Fe(H20):+ 1.4x103 (0.5) 5x103 3.22.107 334 Fe(820)2+ Co(phen)§+ 5.3x103 (1) 1.4x104 1.9x104 337 Co(phen);+ @0120):+ 3O (1) 81 4x103 337 Ru(bpy)§+ Ru(HZO):+ 1.9x109 (1) 5.1x109 3x1012 317 Os(bpy)§+ Ru(HZO)§+ 2.9x108 (1) 7.8x108 9.8x1011 304 Co(phen):+ Ru(HZO)§+ 53 (l) 1.4x102 1.3x105 304 Cr(bpy)§+ v(n20)§+ 4.2x102 (1) 1.1x103 2.6x106 310 Co(phen)§+ «320):+ 4x103 (1) 1.1x104 7.3x105 31o Co(bpy)§+ «1120):+ 1.1x103 ,(2) 2.3x103 2.3x105 407 Ru(NH3)g+ Ru(NH3)§+ 4.3x103 (0.1) 1.9x105 2.2x106 268 Ru(en)§+ Ru(NH3):+ 2.7x104 (0.1) 7.1x1o5 1.6x108 28 Egan):+ Ru(en)§+ 2.8x104 (0.75) 1.3x105 2.5x107 12 Co(en)§+ Co(en)§+ 8x10"5 (1) 3x10-4 1.1x10’4 269 Ru(bpy)§+ Ru(phen)§+ 4.2x108 (0.1) 1.6x109 7.2x108 317 Ru(bpy)§+ Fe(phen)§‘+ 1.8x109 (1) 3.2x109 3.4x1010 317 Ru(bpy)§+ Co(bpy)§+ 2.4x108 (1) 4.3x108 2.1x10lo 408 Ru(bpy)§+ Co(phen):+ 1.4x108 (1) 2.5x108 1.1xlOlo 408 Co(phen):+ Co(phen)§+ 40 (0.1) 1.5x102 5x10”1 409 Co(bpy)§+ Co(bpy)§+ 20 (0.1) 74 5x10’1 409 Co(bpy)§+ Ru(NH3):+ 1.1x104 (0.1) 9.2x104 4.62.106 310 Co(phen);+ Ru(NH3)§+ 1.5x104 (0.1) 1.2x105 1.7x107 310 00(31):;+ Cr(bpy)§+ 35 (0.1) 2.2x102 4.9x103 361 383 those involving considerable inner-shell reorganization around one metal center and none around the other. The free energy driving force can be represented as a displacement of the wells of the free energy curves which are generated in calcula- ting Af and Ar' If both outer- and inner-shell reorganization energies are taken to be quadratic functions of the nuclear coordinates, the energy at the intersection point of the free energy curves, i.e. AG*, can be calculated from AG IX Af (7.7) where the value of the nuclear coordinate X is found by solving Equation 7.8: 2 A X _ _ 2 o f Ar (1 x) + A6 (7.8) Values of AG0 were determined from the formal potentials listed in Table 7.1. Activation free energies for electrochemical reactions were calcu- lated in a similar manner. Solvent reorganization energies for homogeneous reactions were calculated. by ‘using the ellipsoidal cavity' model described. by 374,375 Cannon. The ellipsoidal cavity model allows for interpenetration of reactant coordination spheres, whereas as the conventional two-sphere model of Marcus in principle does not. In light of recent theoretical 14 15 work ’ suggesting that reactant interpenetration is generally re- quired in order to achieve significant coupling of donor and acceptor 384 ENERGY NUCLEAR REACTION COORDINATE Figure 7.1 Schematic representation of the relationships between the nuclear reaction coordinate, forward and reverse reorganization energies, and the classical free energy barrier to electron transfer. 385 electronic orbitals, the Cannon model was considered to be more appro- priate than the two-sphere model. Reorganization energies were calcu- lated according to Equation 7.9: 2 2 2 2 Aout I (Ne r [2131b)(1/sop - 1/E8) S ( )5) (7.9) where e is the electronic charge, r is the distance between the redox centers, 1a and 1b are the semi-major and semi-minor axes of an ellipsoid encompassing the reactants, Sop and 68 are the optical and static dielectric constant of the solvent and S( :3) is the "shape factor.“ This last term depends in a complicated manner on the eccentricity of the ellipsoid, but can be reasonably approximated as 3 (A0) 2‘. 1.19 - 0054 e (7.10) for 0.8§_e:1. The major semiaxis la equals (a1 + a2 + r)/2 with radii a1 and a2 listed in Table 7.1. In the absence of additional information, coor- dination spheres were assumed to interpenetrate such that redox centers were 1.25 X closer than expected from close contact of reactants. Such an estimate is appropriate at least for reactions between aquo 14.15.17 complexes. The minor semiaxis 1b were calculated by taking the volume of the ellipsoid to be equal to that of the two isolated reactant spheres. The values of cop and 68 respectively, at 25°C are 1.78 and 78.3. 386 Solvent reorganization energies for electrode reactions were calculated from13 2 l a — e - e out (Ne Ir)(1/a IIRQ)(1/ op 1/ 8) (7.11) where Re is twice the distance from the redox center to the electrode (i.e. the distance from the charged reactant to its image in the electrode). A value of Re I 78 is obtained if the reaction site is assumed to correspond to the outer Helmholtz plane. Inner-shell reorganization energies were calculated from A $3" - 3fu’m'Ad (7.12) where fM,n+ is the force constant for the symmetrical breathing motions of six identical metal-ligand bonds and Ad is the difference between the equilibrium bond distances for the two oxidation states. Intra- ligand bond vibrations and contortions were ignored since these are anticipated to provide only minor contributions to Ain' values of Ad are listed in Table 7.1. Since structural data are lacking for hexaaquo vanadium(II), the change in the metal-oxygen bond length for the V(H20)2+/2+ couple was estimated from the vanadium-oxygen bond lengthi in related oxide compounds. Beattie, et al. have shown that there is a close correspondence (within circa 0.02 3) between the metal-oxygenIAd values found in oxides and those found in hexa-aquo metal complexes.326 Given the structural similarities, the low spin Fe(bpy g+l2+ g+l2+ and Ru(terpy)§+/Z+ couples were each assumed 3+/2+ )3 . Ru(bpy) , Ru(phen) 387 to require negligible inner-shell reorganization, as for the low spin Fe(phen)g+/2+ couple (Ad I 0.08).376 Force constants were calculated from . 2 2 f AflVinc u (7.13) where c is the velocity of light and u is the reduced mass of the li— gand. For cobalt complexes containing large bidentate ligands, force constants for the cobalt-nitrogen stretch were taken to equal those for hexaammine cobalt (III) and (II). Although values of vin are available for the bidentate ligand complexes suggesting that Equation 7.13 could be employed to calculate force constants, it is not clear what values of the reduced mass should be used since the metal-nitrogen vibration may be partially decoupled from the motion of the ligand as a whole. Nuclear tunneling factors (Tn) can be expressed as AG*(Q) values , thus emphasizing that these represent the difference between the average quantum barrier AG*(T) and the classical Franck-Condon barrier AG*. 19,20,377,378 From Holsteinfis work on small polaron motion, , Sutin, et al. have derived expressions for AG*(T) for reactions without (Equation 9,23 7.14) and with (Equation 7.15) a free energy driving force. The expressions for AG*(T) are: AG*(T) - 1(kT/hv) tanh(hv/AkT) (7.14) and 388 AG*(T) I 0.5AGO’ + (kT/hv){coth(hv/2kT)- [(AG°’)2/Az + csch2(hv/2kT)]1/2 + ( c°’/1)ainh'1[(Ac°'/A)sinh(h»/2kr)]} (7.15) Equations 7.14 and 7.15 were derived for tunneling processes involving a single vibrational mode characterizd by equal force constants and vibrational frequencies in the two oxidation states. Accordingly AG*(T) and refer to only one mode in each calculation. Although AG*(T) values which are calculated assuming equal force constants will certainly differ from those obtained using unequal force constants and frequencies, the difference, AG*(Q), between quantum and classical barriers under simplified conditions should be similar to that obtained when quantum and classical barriers are both calculated more rigorously. Thus, simple single-frequency modes were assumed in calculating AG*(T) and AG*(Q), with the tunneling corrections then being applied to AG* values which had been calculated in a more complicated manner (m m). Tunneling corrections for solvent repolarization were considered to be negligible since v is probably significantly less out than kT/h. Tunneling corrections for each inner-shell mode were calculated separately. Thus the driving force AGO’ appearing in Equation 7.15 was taken in each calculation to be an appropriate fraction “Mm-PAR of the total driving force AGO. This insures that AG*(Q) reaches zero at the same AGO value at which AG* is zero. A point of confusion concerns the use of 1360’ given that Equation 7.15 is 389 'derived using a temperature independent driving force AE°’,. Although the substantial ASo values found for many reactions reflect chiefly the solvation changes accompanying the overall electron transfer process, this entropy change nonetheless would seem to provide a driving force for reorganization of both inner- and outer-shell modes. At least for the present then, A602 is used in place of AEOT. The methods outlined thus far enable rate constants to be calculated using Equation 7.2. However. it is also of interest to com- pare the theoretical predictions concerning activation enthalpies and entropies with the corresponding experimental quantities. Activation entropies (classical limit) were calculated for homogeneous reactions according to: I AS 0 o o calc 0.5AS '0' AG AS [If (7.16) Contributions to AS* arising from the temperature dependence of 1 out were neglected since these amount to 2 J deg-1 11101”1 at most when cal- culated by Marcus’, theory.13 (However, see Section VI. A). Equation 7.16 was derived by assuming equal force constants for inner-shell reorganization in different oxidation states. Nevertheless, this approximation yields little error since the purely thermodynamic term * (0.5 43°) usually provides the dominant contribution to Ascalc' Also, the use of If rather than a "reduced“ reorganization energy nullifies to some extent the errors resulting from the force constant approximation. The required values of A30 were obtained from the thermodynamic reaction 390 entropies 433‘: for the individual half-reactions (Table 7.1). For . . * . . electrochemical reactions ASCalc is given by: * o Ascalc oAsrc (7.17) where a is the electrochemical transfer coefficient and the contribution from Idlf/dT again has been ignored. Classical activation enthalpies were determined from the relation: * t * AHCalc - 11ccalc + 'rAscalc (7.18) where AH* is defined as -R(d1nk/d(l/T)). Nuclear tunneling corrections, An*(Q) and AS*(Q), to activation parameters were obtained by calculating AG*(Q) at various temperatures (Equations 7.14 and 7.15) and noting that - AG*(Q)/ r - AS*(Q), while AG*(Q) + TAS*(Q) I AH*(Q). The resulting tunneling-corrected parameters are designated as AS*(T) and An*(r). 3. Kinetics Formulations and Work Corrections for Experimental W Activation parameters are usually extracted from experimental rate data by assuming a pre-exponential factor of kT/h rather than szn. These "Eyring" (kT/h) activation parameters (designated using superscript daggers) can be converted to "pre-equilibrium” (szn prefactor) activation parameters using: 391 13* - AS #-8 + a 1n(xzvnh/kr) (7.19) and 111* .. AH#+ RT (7.20) thereby enabling straightforward comparisons to be made between theory and experiment. Experimental rate constants for homogeneous electron transfer were corrected for the electrostatic work of precursor formation by way of the Debye-Huckel treatment. This approach is embodied in Equation 7.21: log k corr I logk + zAzBezN/2.303RT ssr(l+8r111/2) (7.21) where 2A and 2B are the reactant charge numbers,8»is the Debye-Huckel parameter,11is the ionic strength and Rear: is the rate constant which is expected in the absence of electrostatic interactions. Since several kinetics studies have indicated that activation enthalpies decrease systematically with increasing ionic strength while activation entropies remain essentially unchanged,28’379-381 it was deemed appropriate to incorporate work corrections wholly in AH*. Electrochemical rates were corrected for electrostatic diffuse double-layer effects by using the Gouy-ChapmanISternIFrumkin theory as described in Sections 1.0 and 4.0. 392 C. Results Calculated rate constants are compared in Figures 7.2 and 7.3 with work-corrected experimental rate constants. The reactions are separated into two groups: those involving a pair of reactants possessing iden- tical ligand compositions (Figure 7.2) and those involving reactants possessing dissimilar ligand compositions (Figure 7.3). In order to compare electrochemical and homogeneous reactions on a common basis, rate constants for the former (kE ) were converted to second-order calc E rate constants by multiplying k values by the ratio of statistical calc factors for the two types of reactions, i.e. Rz/Gr. values of kcalc and kcor r are listed in Tables 7.2 and 7.3 together with observed experimental rate constants. values of AC*(T), AG*(Q), AC°,Af, E2 and Vn are given in Table 7.4. The calculated rate constants plotted in Figures 7.2 and 7.3 were obtained using individual force constants and tunneling corrections for the inner-shell, with the Cannon model being used to calculate the outer-shell barrier for the homogeneous reactions. The use of individual force constants yielded rates that were typically three-fold, but in a few cases as much as twenty-fold, smaller than calculated using reduced force constants. Allowing for reactant interpenetration usually decreased. .AG*(T). Nuclear tunneling corrections increased the calculated rate constants. Values of In range from 1 to 17, the largest values being found for low driving force reactions requiring substantial inner-shell reorganization. values of l 1 K0 range from 0.13 M- A for reactions between aquo complexes to 0.9 M- for reactions between polypyridine complexes. 393 l I T I // '0’ /18 w 3’. / 6 )9 / 405° / / / g )1 A: 7 5- // . 359:9 _. t / I'oo 5 x5 / / I0 4| / O .5 3 3 g / ‘ o I? / // 0e / _ / I / o / / 36 / 5 L l l l -5 0 5 IO RMIMT) Figure 7.2 The log of the work-corrected experimental rate parameter kcorr plotted against the log of the corresponding theoretical parameter k(T) for electron transfer between reactants possessing identical ligand compositions. Key: (C» aquo complexes: (A) polvpvridine complexes: (O) ammine complexes. Reactions and rate constants listed in Table 7.2. 394 I no» ‘ 35 S _ :5 5" !. 3 5 _ .1 0’ 3 -5 1 l l '5 0 5 l0 log k(T) and Iog[K;km/sr] Figure 7.3 The log of the work-corrected experimental rate parameters 0 . kcorr and KAkcorr/ér plotted against the log of the corresponding theo retical parameters k(T) and K:k(T)/6r for electron—transfer between re- actants possessing dissimilar ligand compositions. Key to cross-reactions: GB) aquo-ammine; (A) aquo-polypyridine: (A) ammine-polypyridine; (-—9 electrochemical. Reactions and rate constants are listed in Tables 7.2 and 7.3. 46 47 48 49 50 51 Table 7.3. 395 Experimental and Theoretical Rate Constants for Electrochemical Electron-Transfer Reactions k (kJ-mol-l) (cm 5.1) AG° Reactant Ru(NHB) 2" o 410 at mercury Ru(H20)2+ o 313 at mercury Fe(320)2+ -20 313 at mercury _30 -40 —so «1120)? +30 313 at mercury +20 +10 0 -1o v MM Aavsoa Novgoqn “a we .oal coauummm AsmacfioaouV s.N «News 399 x . «Hod m m Haoaxm x Haoa 0 x maoa H x ~H0H m x O NH0H 0 m NHOqu x Naoa m x ~H0H a x ~H0H 0 x «Hod 0 x O NH0H 0 m Naoaxo n As-mv n— > 00.0 00.0 m0.0 mN.0 mm.0 NN.0 mma.0 00.0 00.0 00.0 00.0 00.0 0<.0 A .mTE H < oz m.ma n.0H mm mN AN m.0m Hq mm mm Na 0 a Ayvswc A -Hoa was 0.0 0.m H.0 m.0 m.~ q.~ 0 -Hoa was A00x0 a« monouoa um «« no- a..- 2. o missions 3 N12. w. 08 we. o8 w. 9: w. 05 m. . H1 we so owuH H1 n so AH1H H1 we so AH1H H1 we no AH1H H1 we so AH1Hoa axe mucmuommm HHOU Aav«ma on«m< «ma «we .mq .ua .mooHuomom uommcmua1couuoon HmoHaonoouuoon you moouom wcH>Hu0 oHam:0003ho:H 0mm moHaouumm aoHuo>Huo< HmoHuouoonk 0cm HmumoaHuonxm .0.m oHan 409 m.m1 m.m m.H1 n.H1 «oH1 +MAoN=van +MAsanvam aH HN HH «.mm mm m.H +onN=V> +MA=mvoo RH mN H.« «1 m.N Nm1 +onN=v> +MAN=zvae 0H NN N.H m.NH mH NH1 +MAamvmm +MAoN=vmn NH m.mN m.« m.o m.mH «m1 +wAmmzvsm +MAonvma HH mN a.c «H m.oH Nm1 +onN=v> +MAoN=vmm a m.«m N.o m.mN m.mN H«1 +onvaam +MAoN=vmm N «N H.N an n.mm NHN1 +onvauu +MAoN=Voo m m.mm m.HH m.Nm m.mm ma1 +onN=me +onvaoo « m.ooH N.mH m.m« m.a« _o +MAonvoo +MAonvoo m mm o.m mm an o +onN=voa +onN=Voa N m.mo N.m Hm m.Nm o +onva> +MAoN=v> H Asv«=< onsm<1 sea em< .=< namuusuan osmeon .mnoHuomom uommmmu91mouuoon msommmwoaom “00 moouom me>HHO AHIHoE 0x0 mnHozumm 0Hammmvoauone 0mm moHnHmnucm cOHum>Huo< HmoHuouoonh 0mm HoumoaHuomxm .50 mHan. 410 as o.N m« m.o« o« +MAsapvno +MAcmvou m« m.oN m.e mm Nm o +MAsanvoo +MANaavoo N« n.oN m.o m.oN m.HN o +mAcoeavoo +MAcmeavoo H« m.oH o m.Hm NN o +wAamnavsm +MANanv=m Nm «m «H «.mm n.0m o +mAaovou +MAamvoo on NH «.o m.« OH mH1 +wAmmzv=m +MAamV=m «m om N.o m.om m« o oAmzzvsm oAmmzvae mm +N is aN o.N OH mH HN1 +onva> +MAsanvoo Nm m.oN m.o 0H oH mN1 +onva> +MAcmzavoo Hm m.n1 o m.N1 m.N1 +onva=m +MAsanvmo wN m.o+ «.N m.«1 m.m1 mm1 +onvamm +MANanvom NN m.H o.« o m.o 0N1 +onvamm +MAcmnavoa NN «.N1 m.m m.HH1 m.HH1 «oH1 +onN=0mm mAsanmuVam HN m.m1 w.m m.«1 m1 «0H1 +onvama +MAamnavam ON Aevsm< onm<1 «:< +m< om< ucmuomvou ucmvon AvosmHucoov N.N oHan 411 N N.H «61 m.«1 81 +MA5e38 M2303 o« no NH «H1 NH1 31 +MA2£8 +MAE£§ mm QAN 9m m.«H NH 0N1 +wAmmzv§ $2238 «« AHvsm< A0001 «mq +00 o=< ummuozvou uomvon AsmacHuaoUV N.N 8H e3. 412 00 N.N1 0.N0 0.N0 H 001 0.H0 H.ml m0 00 HH 00! vNQH.um 0.00 0.01 00 m0 HN O0l +MAON=vHU 0.00 n.01 00 H0 H0 Oml 00 0.01 0h 0m H0 ONI 0.0m 0.01 0N H0 H0 0H1 hunches um mam 37 «m 3 Ho o +m Aonto 0.00 0.01 0.00 00 00 0H1 Assumes on No «1 m. H... «o 3 o +m 8N5> humonoa um m.NN o 0A032v=¢ +m uuoo Aavsmq Hovsm< «00 £50 000 000 munmuomom .mcoHuomom ummmmmue1aouuoon Honaonoouuoon How AH Hos 0x0 moouom waH>Hu0 maHmnucm oHamnavoanone 0cm moHnHmsumu moHum>Huo< HmoHuouoonh 0cm HmucoaHnomxm .0.N oHan 413 9% «.m1 Hm Hm HN o«1 n no N m1 m on m on Hm 81 0me 95 a... 2 H31 N« N« H« 8. +onN£5 «m 2.0.32 A332 .Huommq «:0 .3 .3 muamuummm AvoamHuooov 0.N oHan 414 D. Discussion The overall impression that is conveyed by the rate comparisons in Figures 7.2 and 7.3 is that the theory of weak-overlap electron transfer is remarkably successful in describing the energetics of such reactions. The experimental rate constants do tend to be smaller than the calcu- lated values. This might be taken as an indication that nonadiabaticity prevails in most reactions even at the closest approach distance. Nevertheless, it is worthwhile to search further for explanations of rate disparities. A useful approach is to examine the reactivity of metal complexes as a function of ligand and electronic structure. Reactions of aquo, ammine and electrode co-reactants are examined in Figure 7.6 by plotting the difference between log k and log k(T) against a normalized free corr energy driving force coordinate. Although the data for the homogeneous reactions are scattered, the overall indication is that the reaction rates increasingly deviate from theory as the driving force increases. This trend is reminiscent of earlier observations based on relative rate comparisons via the Marcus cross relation.3m’311 (See Section VI. B). However. explanations in terms of the shortcomings of the cross relation are inapplicable here, since the various approximations inherent in that approach (cancellation of electrostatic work terms, additivity of outer-shell reorganization energies, etc.) are not made in the present analysis. The driving force dependence of the breakdown of theory suggests that the explanation lies in the calculation of Franck-Condon barriers, rather than in neglect of work terms, nonadiabaticity or other pre- 415 l I l 5 30 H S% 4 o 9 12 o O ‘8 2.. .1 .5. x T; v 5 8 I’- — Or- 15 H 2 o 111 l L O I 2 3 - 8m0/(Af‘b'lr) Figure 7.6. Log [k(T)/kcorr] for reactions involving aquo complexes 0 . . plotted against the reduced driving force AC /(lf+ir). Key to reaction type as in Figures 7.3 and 7.4. 416 exponential factors. Interestingly. the apparent breakdowns for homo- geneous reactions are paralleled in the electrochemical oxidations of “320):+ and Cr(H20)§+. but not in the corresponding electrochemical reductions. Nearly all of the homogeneous reactions included in Figure 7.6 also involve aquo oxidations. A reasonable hypothesis is that the driving force dependent rate discrepancies for both the homogeneous and electrochemical reactions arise because the rate calculations over- estimate the extent to which the activation free energy for the aquo complex is diminished by increasing the driving force. A speculative explanation for this centers on hydrogen bonding between aquo ligands and the surrounding solvent. It is usually assumed that Aout possesses the same value for forward and reverse reactions. Nevertheless there is ample evidence of enhanced hydrogen bonding in the M(EZO):+ state com- 55,222 3+ This suggests that A could . 2+ pared with the M(HZO)6 state. out exceed Aoutzfi which in turn would yield an even greater asymetry between aquo oxidation and reduction reactions than is expected from inner-shell considerations alone. A possibility which cannot be dismissed is that the driving force dependence of log [(kcorr/HTH represents merely a coincidental ordering of experimental and calculational errors, at least for the homogeneous rate data. Indeed, the calculated rate constants could be in error by as much as a factor of twenty in some cases due to errors in estimating An and therefore xin' Such uncertainties emphasize the need for accurate measures of symmetrical metal-ligand stretching frequencies in addition to data concerning metal-ligand bond lengths. 417 3+/2+ 6 since it has often been suggested that these are somehow anomalous. The The cross reactions involving Co(HZO) are of special interest rate comparisons summarized in Figures 7.2, 7.3 and 7.6 indicate that while such cross-reaction rates do indeed deviate frmm the predictions of theory, these deviations are comparable to those found for other reactions. The Co(HZO)2+’2+ self-exchange rate however, exhibits enormous deviations from theory. Endicott and coworkers have suggested that the hexaaquo cobalt (III)I(II) self-exchange follows a catalytic inner-sphere pathway, a conclusion which is strongly supported by the present calculations of outer-sphere rates.384 Curiously, the activa- tion entropy for this reaction is typical of that found for genuine outer-sphere reactions (Figure 7.4), suggesting that the magnitude of AS* is not particularly diagnostic of the reaction mechanism. Reactions between aquo and polypyridine complexes form an interesting group. On average the cross reactions involving low-spin polypyridine complexes with aquos are nearly four orders of magnitude . slower than predicted (Figure 7.7) while the rates of corresponding reactions of pairs of aquo complexes and pairs of polypyridine complexes evidently agree more closely with theory. The discrepancies between kcorr and k(T) appear to be independent (or nearly so) of the reduced driving force for the reaction. These observations suggest that some type of barrier to precursor formation, perhaps arising from an energetically unfavorable interaction between aquo and polypyridine ligands, might be the source of the rate discrepancies. Alternative explanations of rate disparities in terms of miscalculation or neglect of factors contributing to nuclear reorgani- 418 23 13 _ 19. 20.21 _ a a 32 29 22 a log 111(T)/11.°,,1 1 1 1 1 1 O I 2 3 -8 AGOI (Ag Ar) Figure 7.7. Log [k(T)/Reorr] for reactions between polypyridine and aquo complexes plotted against the reduced driving force AGO/(Af+lr). Key: (A) low-spin polypyridines with aquos; CV) high-spin polvpyridines with aquos. 419 zation energies appear to be untenable. The effects of such errors should be attenuated with increasing exoergonicity, yet rate discrep- ancies persist even in the 8110120):+ - Ru(bpy);+ cross reaction for which AGO approaches if. The behavior of aquo complexes with polypyridine reactants con- trasts with the behavior of aquo complexes with other coreactants. One might speculate that some of the pecularities of 110120)?”2+ reactivity, which are tentatively attributed here to hydrogen bonding interactions with the solvent, are somehow circumwented when the aquo reactant is placed in close proximity to a polypyridine coreactant. Interestingly solvent "structure breaking“ capabilities have often been attributed to the latter.306 The activation parameters for the aqua-polypyridine reactions are characterized by AS* values which differ from theory by some 85 J deg-1 mol-l, a discrepancy similar to that found for a variety of other cross * reactions (Figure 7.4). In contrast, AH values are quite similar to corr the predicted values. Such enthalpic behavior.distinguishes these cross values that are * 01': reactions from most others, which instead exhibit AH.c smaller than expected (Figure 7.5). Evidently, AH* rather than AS* is the parameter signaling a typical rate behavior. It is perhaps not use- ful to speculate concerning the details of the postulated work term for aqua-polypyridine reactions, beyond suggesting that it is partly enthalpic. . . . 3+/2+ Since cross reactions between aquo couples and either Co(bpy)3 or Co(phen)§+/2+ agree more closely with theory than do the corre- sponding reactions involving low-spin polypyridine couples, it is worth- 420 while to ask whether these differences are related to the differences in electronic structure between the cobalt couples and the others. Although this is an intriguing possibility, the most likely explanation is that the reactivity differences are due to miscalculations of metal-ligand bond reorganization energies for the Co(phen)§+/2+ and Co(bpy)3+/2+ 3 couples. Substantial calculational errors are certainly possible for these couples given the tentative nature of the force constant estimates and the 18r8e uncertainties (10.027 8) in the EXAFS estimate of Ad. For reactions requiring very large adjustments of bond lengths even small uncertainties in Ad become significant in the rate calculations. This suggestion also accounts for the otherwise puzzling result that kcorr exceeds k(T) for the Co(bpy)g+/2+ and Co(phen)g+/2+ self-exchanges. Note that catalytic inner-sphere pathways which might cause kcorr to exceed the calculated outer-sphere rate constant are not available since both the Co(III) and Co(II) centers are substitutionally inert. “Cross reactions“ between metal complexes and electrode surfaces form a distinctive set of electronrtransfer reactions, one in which the thermodynamic driving force can be varied without altering the chemical identity of the reactants. Complexes exhibit different degrees of outer-sphere reactivity with different electrode co-reactants, even after correcting rate constants for electrostatic double-layer effects. Such differences are intriguing since they are unexpected from electron transfer theory. Despite the differences in kco values, the varia- rr tions of rate constants with driving force are essentially the same for the reduction of Cr(HZO):+ (and other reactants) at different surfaces. Liu has suggested that differences in kcorr can be understood in terms 421 of the degree of hydrophilicity and corresponding interfacial solvent ordering that is exhibited by each metal surface.136 The least hydro- philic surface, mercury. appears to correspond most closely to an ideal “weakly interacting” coreactant, while the most hydrophilic surface, gallium. strays furthest from ideal behavior. Activation parameters are markedly more sensitive to electrode composition that are the kcorr values. The electrostatic effects of double-layer structure an electrochemical kinetics parameters are better understood than the corresponding salt effects on homogeneous kinetics. Thus it can be claimed fairly confidently that the variations in activation parameters, as well as rates, represent something other than uncertainties in electrostatic work terms. The apparent importance of electrode-solvent interactions suggests that specific solvent interactions with homo- geneous co-resctants, most notably polypyridine complexes, may also bear some relation to electronrtransfer reactivity. There are too few examples of each of the other groups of cross reactions to be able to isolate ligand-specific reactivity patterns. A general observation however is that a rather less optimistic view of the success of electronetransfer theory emerges when cross reactions are included in rate comparisons. Comparisons between calculated and experimental activation parameters, which have been alluded to in the foregoing discussion, form a more demanding test of theory than do the rate comparisons. Overall the results summarized in Figures 7.4 and 7.5 are nothing short of re- markable. Clearly. it is possible to account quite adequately for 422 variations in AB? and especially, AS* as the thermodynamic and struc- tural properties of the reactant are varied. On an absolute level however. the agreement between experiment and theory is less satisfactory. we noted above that a discrepancy of 85 1 1 25 J‘deg- mol-1 is observed between A3* and As:alc for nearly all corr the homogeneous reactions. One explanation is that outer-sphere elec- tron transfer reactions are in most cases strongly nonadiabatic, such that the value of :31 is about 10‘s. Besides being inconsistent with the few 52 inigig studies of electronic coupling between outer-sphere 14,15 reactants, such a value is simply too small to account for the observed rate constants. Various other possibilities were considered in a recent examina- tion of the effects of reactant-solvent hydrogen bonding on activation entropies (Section V. A). There it was concluded that the differences between As:orr and As:alc almost certainly cannot arise from the tem- perature dependence of the Franck-Condom barrier, i.e. AG*(T), but instead must represent a work term associated with precursor formation. Friedman has arrived at a similar conclusion on the basis of his studies of ion-pair correlation functions for aquo complexes.17 Specifically, the negative activation entropy for the Fe(H20)2+/2+ self-exchange is assigned in his work to coulombic repulsion between the charged reac- tants. A significant finding is that the entropic component of such electrostatic interactions is expected to be nearly independent of ionic strength. The common feature of the set of homogeneous reactions exam- ined here is that each involves a reactant of charge 2+ and another of charge 3+. Provided that electrostatic interactions are controlled 423 essentially entirely by the charges of the reactants, with the reactant size and ligand composition being of little importance, such inter- actions provide a very plausible explanation. various other factors such as steric requirements, marginal nonadiabiticity, variations of Kel with temperature. etc. likely contribute, at least to a minor extent, to the observed discrepancies between experiment and theory. Returning to the adiabaticity question, an unexpected finding is that the 88* values for reactions involving cobalt centers are no more discrepant than for other reactions. For the cobalt couples electron transfer is a formally spin-forbidden process. Jortner and co- 385 workers have noted that the spin restriction an be partially overcome through spin-orbit coupling, leading to an electron transfer probability in the case of Co(NH3)2+’2+ self-exchange that is 10-4 smaller than expected in the absence of spin restrictions. Since evidence for such spin-related nonadiabaticity is wholly lacking in the rate and activa- tion entropy comparisons (note also the conclusions for Section VI. C), one suspects either that spin-orbit coupling is a more successful reac- tion scheme than previously believed or that another mechanism is available to circumvent the spin restriction. Offsetting the unfavorable experimental activation entropies are activation enthalpies which tend to be more favorable than predicted. Overall, the differences between1Aflzorr and1AR*(T) may possibly have the same origin as the AS*(T) - As:orr differences, namely. electrostatic interactions that differ from those assumed in making work term correc- tions. The larger scatter in AH* values compared with AS* values is not unexpected, since calculations of activation enthalpies are more sensi- 424 tive to errors in estimates of reactant structural parameters. However, scatter in the (18* data could additionally result if breakdowns in theory arising from inter-reactant and reactant-solvent interactions, as well as departures from purely outer-sphere reaction pathways, are manifested chiefly as enthalpic effects, as has been suggested above. E. Conclusions In a global sense electron-transfer theory is reasonably successful in predicting the rates of self-exchange and cross reactions as well as those for electrochemical reactions. The lingering discrep- ancies of 10 to 104 -fold seem to arise chiefly from specific, non- electrostatic reactant-reactant and reactant-solvent interactions. Cross reactions between aquo and polypyridine couples are the best example of the former, while electrochemical reactions are the strongest example of the latter. Nonadiabaticity evidently is not exceptionally important in reactions requiring formally spin-forbidden electronic transitions, although in general nonadiabaticity certainly could be significant. The variations of activation parameters with thermodynamic and structural factors generally are well described by theory. However, the theoretical estimates of AH" and 118* are substantially different to the experimental values. Such errors, at least in homogeneous reactions, probably represent mainly miscalculations of electrostatic work terms rather than failures of electron-transfer theory itself. 425 CHAPTER VIII CONCLUSIONS AND SUGGESTIONS FOR FURTHER WORK A. Conclusions Differential capacitance experiments indicate that silver is capable of adsorbing several simple inorganic anions to the extent of a monolayer at the most positive accessible potentials. Electrochemical toughening which is required in order to observe SERS signals from adsorbed ions was found to produce only minor changes in the average surface concentrations of such anions at silver. Underpotential deposition of a single atomic layer of lead on to a silver electrode was sufficient to transform the double-layer structure and adsorption properties of the surface essentially to those of a bulk lead electrode. Thus, adsorption is greatly diminished at the UPD lead/silver. as well as UPD thallium/silver surfaces in comparison to unmodified silver. The differences in anionic adsorption properties between these three surfaces as well as mercury appear to be related to differences in the degree hydrophilicity of the metal electrodes. In order to evaluate electrochemical kinetics at a molecular level and to correlate electron-transfer reactivity in different environments an "encounter preequilibrium" treatment was refined and further devel- oped. This treatment provides a more satisfactory theoretical descrip- tion of redox processes than is obtained from the conventional colli- 426 sional model. In particular, a clearer picture of the physical signifiance of the frequency factor emerges from the former. The preequilibrium treatment was applied to the problem of the influence of the metal surface composition on redox reactivity. Rates of inner-sphere electron transfer reactions at silver are diminished by three to five orders of magnitude when the surface is modified by underpotentially depositing a monolayer of lead or thallium on silver. An analysis on the basis of the encounter pre-equilibrium model indicates that the reactivity differences are associated with a change in reaction mechanism from inner- to outer-sphere and an accompanying decrease in the precursor stability constant. Surprisingly, rate constants for the elementary electron-transfer step for several chromium(III) reductions are closely similar at four different electrode surfaces. A careful consideration of entropic driving forces for intra- molecular electron-transfer indicates that, contrary to common belief, such reactions typically involve a significant Franck-Condon activation entropy. This complicates the analysis of frequency factors from the point of view of monitoring nonadiabaticy. Nevertheless when carried through, such an analysis can account for peculiarities such as positive activation entropies and the marked sensitivity of apparent frequency factors to nonrbridging ligands. From an evaluation of rate responses to systematic alterations of double-layer structure together with rate comparisons for parallel inner- and outer-sphere reaction pathways it is concluded that chromium(III) aquo reductions are mildly nonadiabatic at the mercury- 427 aqueous interface. The corresponding ammine reductions are decidely adiabatic. A generalized comparison of electrochemical and homogeneous redox reactions of aquo complexes provides good evidence that driving force dependent breakdowns of the relative electronetransfer theory for the two data sets have a common origin. Subsequent absolute calculations indicated that such breakdowns result in part from differences in force constants for inner-shell reorganization between different oxidation states. Comparisons between self-exchange and cross reactions as well as between homogeneous and electrochemical exchange reactions provide substantial evidence that the homogeneous self-exchange reaction of the hexaaquo iron(III)/(II) couple probably follows an extraordinary reaction pathway in comparison to other reactions involving aquo com- plexes. It was speculated that this pathway might involve water as a bridging ligand. Since the Fe(H20)2+l2+ self exchange has served as the prototype reaction for evaluating outer-sphere electronetransfer theories, information concerning the reaction mechanism should be of considerable interest. Absolute theoretical estimates of rate constants for nearly sixty outer-sphere homogeneous and electrochemical reactions exhibit tolerably good agreement with the experimental rate constants. However. the agreement between theoretical and experimental activation enthalpies and entropies is considerably worse. Nevertheless, this poor agreement appears to be due to the inadequecies of the treatment of elecrostatic work terms rather than the electron-transfer theory itself. 428 1!... W m 2.113322; M There are a number of experiments involving ionic adsorption which might be expected to yield interesting surface chemistry. Because of the enormous range of positive electrode charge available at silver in adsorbing electrolytes diffuse-layer potentials should be an important factor in anionic specific adsorption. The influence of the diffuse layer could be investigated by examining specific adsorption over a range of ionic strengths. One suspects that diffuse-layer effects at very large electrode charges might be coupled to electrosorption valencies, Frumkin g parameters and other isotherm parameters in peculiar and interesting ways. The simultaneous ionic adsorption analysis scheme outlined in Section III. C could be tested by monitoring specific adsorption from chloride-bromide-perchlorate or chloride-aside-perchlorate mixed elec- trolytes at silver. If successful, such experiments would be useful for establishing more satisfactorily the connections between SERS intensi- ties and average surface concentrations of Raman scatterers. A fault of the experiments described in Section III. A is that the electrode was roughened in one electrolyte, soaked in a second solution and subjected to capacitance measurements in a third electrolyte. All of these steps could be performed with a single electrolyte solution, thereby parallel- ing more closely the experimental protocol employed in SERS experiments, if the proposed analysis scheme were used in place of the conventional Hurwitz-Parsons‘method. Another sequence of experiments might focus on the thermodynamic properties of metal electrodes which are covered only partially by an 429 underpotentially deposited metal. A nonlinear relationship was found between the degree of surface coverage of silver by UPD lead, and the magnitude of the differential double-layer capacitance. Such nonlinear relationships might be observed with regard to the ionic adsorption pro- perties of such surfaces as well and would have interesting consequences in connection with theories of adsorption. Additional insights concerning electron-transfer energetics at underpotentially deposited lead and thallium electrode surfaces might be gained if suitable inner-sphere reactions could be found. Chromium(III) thiophene reductions represent one possibility. An investigation of solvent isotope effects on outer-sphere reductions at UPD lead/silver and UPD thallium/silver might provide clues as to the importance of specific interactions between reactants and the solvent inner-layer since hydrogen-bonding interactions generally are stronger in D20 than H20. The relative electron-transfer theory could be used to elucidate various features of the electrochemical reduction kinetics of chromium(III) complexes if corresponding data could be gathered for homogeneous outer-sphere reactions of such complexes with a common reductant such as Os(NH3):+ or Co(bpy)3+. An improtant conclusion from the study 'of absolute electron- transfer reactivity (Chapter VII) is that the usual treatment of elec- trostatic work terms is inadequate for homogeneous redox reactions. In order to develop a more satisfactory treatment it would be useful to measure activation parameters over a range of ionic strengths for a set 430 of reactions involving a number of different reactant charge combina- tions. .An improved description of electrostatic effects on kinetics parameters would be extremely valuable if it enabled such effects to be separated from additional subtle factors (e.g. nonadiabaticity) which may well be masked in many circumstances by the current (inadequate) treatment. Finally, it was noted that the absolute theory of electron transfer predicts that negative activation enthalpies will be observed under certain conditions. Indeed negative values have been found for the Fe(H20)§+-Fe(bpy)g+ and other cross reactions. It would be worth- while to search out electrochemical rections which might exhibit such behavior. The chief requirement according to theory is that the entropy driving force (-TAS°) for electron transfer must exceed in absolute value the intrinsic reorganization energy (If). Oxidations of ammine complexes in nonaqueous solvents such as acetonitrile involve enormous entropy changes (circa 200 J deg-1 mol-l) or more; (see Section V. C). The entropy driving force effects might be sufficiently large to reveal not only negative ideal activation enthalpies, but also an “inverted“ enthalpy region where AH* increases with increasing exothermicity. An enthalpic inverted region is expected under extreme circumstances from theoretical considerations. Rate behavior in this region might be in- fluenced by anomalous nuclear tunneling effects or other unusual fac- tors. In any case this problem appears to be worthy of additional cogitation, as well as experimental investigation. 431 APPENDIX I Reaction Entropies for Transition-Metal Redox Couples in Various Solvents This appendix consists of a listing of previously unreported As:c data which were obtained in connection with the work described in Section V. C. Table A.1. Reaction Entropies (J deg- 432 1 Couples in Various Solvents. Redox Couple )3+/2+ 3 6 3+/2+ Ru(NH3)6 3+/2+ 3 Ru(N83)5pz Ru(NH Ru(en) 3+/2+ 2+ Ru(NH3)5pz3+/ 3+/2+ Ru(NH3)4bpy 3+/2+ Ru(NH3)4phen 3+/2+ Ru(NH3)4phen Ru(NH3)4phen3+/2+ 3+/2+ Ru(NHS)2(bpy)2 3+/2+ 2 3+/2+ Ru(NH3)2(bPY)2 3+/2+ Ru(NH3)2(bpy)2 Ru(bpy)§+/2+ Ru(bpy)§+l+ Ru(NH3)2(bPY) Ru(bpy)}:l0 Cr(bpy)g+/2+ Cr(bpy)§+l+ Cr(bpy);/o Co(bpy)§+/2+ Solventa acetonitrile acetone formamide nitromethane propylene carbonate acetonitrile nitromethane dimethylsulfoxide propylene carbonate nitromethane acetonitrile dimethylsulfoxide propylene carbonate acetonitrile acetonitrile acetonitrile acetonitrile acetonitrile acetonitrile acetonitrile mol-1)for Transition-Metal Redox Asrc 185 200 90 165 155 155 120 125 150 115 130 110 135 115 70 25 105 65 20 190 433 Table A.l (continued) I'C Redox couple Solvent ASO Co(bpy) :14" acetonitrile 60 Co(sep )3” 2+ acetonitrile 210 Co (EFME—Oxo sar-H) 2+“ water 30 Co(EFME-Oxosar-H) 2+, + N-methylformamide 9O " acetonitrile 165 " formamide 75 " ' dimethylformamide 175 " dime thylsulfoxide 160b " propylene carbonate 150b " methano l 16 5b " ni trome thane l40b 1' ethano 1 1051’ a. 0.1 .11 RPF6 supporting electrolyte b. measured by Dr. Peter Lay 434 APPENDIX II Negative Activation Enthalpies In comparing experimental and theoretically calculated activation parameters (Chapter VII) it was noticed that in a few instances negative activation enthalpies have been observed for homogeneous outer-sphere 329,359 reactions. Marcus and Sutin have pointed out that this peculiar result is actually expected under certain conditions from.the relative 246 In connection with the work described in electronrtransfer theory. Chapter VII it was of interest to ascertain under what conditions the absolute electronrtransfer theory would predict negative values of AH*. The problem is somewhat simplified by assuming equal inner-shell force constants for the oxidized and reduced states of each reactant, and a value of the intrinsic activation entropy equal to zero. It is found that negative 1111*~ values can be expected within the so-called "normal" free energy region (AG°