Sie sind auf Seite 1von 58

Chem. Rev.

1998, 98, 113-170 113

Homogeneous-Phase Electron-Transfer Reactions of Polyoxometalates†


I. A. Weinstock
USDA Forest Service, Forest Products Laboratory, One Gifford Pinchot Drive, Madison, Wisconsin 53705,
and the Department of Chemistry, Emory University, 1515 Pierce Drive, Atlanta, Geogia 30322

Received September 23, 1997 [Revised Manuscript Received November 12, 1997)

Contents VII. Reduction of Inorganic Oxidants by Reduced 156


POM Anions
I. Introduction 113 A. Introduction 156
II. Outer-Sphere Electron-Transfer Reactions 115 B. Hydrogen Cations 156
A Debye-Smoluchowski and Marcus 115 C. Nitrite 157
B. Marcus Equation 116 D. Nitronium Ion 158
1. Applications 116 E. Peroxodisulfate 159
2. Theoretical Calculation of Outer-Sphere 117 F. Chlorate and Periodate 159
Electron-Transfer Rate Constants G. Permanganate 160
3. Outer-Sphere Electron Self-Exchange 117
VIII. Reduction of Dioxygen by Reduced POM Anions 161
4. The Marcus Cross-Relation 117
A. Introduction 161
C. Ion Pairing between Electrolyte Ions and 118
B. Single-Electron Reduction of O 2 161
Electron Donors or Acceptors •–
119 C. The O 2/O2 Self-Exchange Reaction 161
D. Outer-Sphere Electron Transfer within Stable
Donor-Acceptor Complexes D. Reduction of O 2 by Reduced Isopoly- and 162
121 Heteropolytungstates
Ill. Electron Self-Exchange between POM Anions
121 E. Reduction of O 2 by d-Electron-Containing, 164
A. Keggin Tungstocobaltate Anions Transition-Metal-Substituted
B. Keggin and Wells-Dawson Metallophosphate 122 Heteropolytungstates
Anions F. Reduction of O 2 by Reduced 166
C. Reevaluation of the Reorganization Energy of 125 Vanadomolybdophosphates
the [CoIIIW12O40]5–/[COIIW12O40]6– Couple IX. Acknowledgments 167
D. One-Electron Oxidation of [CuIW12O40]7–to 129
X. References 167
[Cu IIW12O40]6–
IV. Oxidation of Organic Electron Donors by POM 129
Anions I. Introduction
A. Alkylaromatic Compounds 129
B. Cyclohexadienes 136 Many isopoly- and heteropolyoxometalate (POM)
C. Organic Acids, Esters, Aldehydes, and 138 anions are ideal candidates for the observation of
Ketones homogeneous-phase outer-sphere electron-transfer
D. Aromatic and Aliphatic Alcohols 142 processes and for comparison of experimentally de-
E. Carbon-Centered Organic Radicals 143 termined rate constants with those predicted by the
145 theoretical models of Marcus, Sutin, and others. At
F. Thiols and Organic Sulfides
the same time, where one or more addendum
V. Oxidation of Inorganic Electron Donors by POM 148
Anions
atoms-usually V(v), Mo(VI), or W(VI)–of the POM
cluster have been substituted by d-electron-contain-
A. Introduction 148
ing first- or second-row transition-metal cations, more
B. Inorganic Electron Donors 150 rapid inner-sphere electron-transfer pathways are
C. Hydrogen Sulfide 150 often accessible. The purpose of the present article
D. Luminescence Quenching of Electronically 152 is to provide a comprehensive and critical analysis
Excited [Ru(bpy) 3]2+ Cations of published data concerning the mechanisms of
VI. Reduction of Halogenated Alkanes by Reduced 153 homogeneous-phase electron-transfer reactions be-
POM Anions tween POM anion pairs, and between POM anions
A. Carbon Tetrabromide 153 and organic or inorganic electron donors or acceptors.
B. Halomethanes 154 The elegance and utility of Marcus’s theory of
C. Carbon Tetrachloride, Halomethanes, and 155 outer-sphere electron transfer were recognized in
Haloalkanes 1992 with a Nobel Prize in chemistry, and its histori-
cal development was outlined in his Nobel Lecture.1
* Visiting scientist at Emory University, 1996-1998. Permanent Shortly after publication of Marcus’s seminal papers
address: USDA Forest Service, Forest Products Laboratory, One on this topic,2–4 it was recognized that POM anions
Gifford Pinchot Drive, Madison, WI 53705. were likely to exchange electrons via an outer-sphere

Dedicated to Dr. J. Hodge Markgraf, Ebenezer Fitch Professor
of Chemistry at Williams College, on the occasion of his retirement. mechanism: (1) As a class, POM anions possess
S0009-2665(97)00341-5 This article not subject to U.S. Copyright. Published 1998 by the American Chemical Society
Published on Web 01/22/1996
114 Chemical Reviews, 1998, Vol. 98, No. 1 Weinstock

as a well-defined probe for determining the nature


of outer-sphere oxidations of alkyl aromatic hydro-
carbons.7 His aim was to determine whether such
processes were feasible.
An additional rationale for using CoIIIW125– was the
expectation that reduction of tetrahedrally coordi-
nated Co(III) ions (ground-state electronic configu-
ration e3t23, 5E) to tetrahedral Co(II) (e4t23, 4A) would
occur rapidly in the absence of electronic excitation.
In contrast, addition of an electron to an eg orbital of
low spin octahedral Co(III) (t26e0, 1A) is formally a
III 5–
spin-forbidden process. Hence Co W12 anions
were expected to be more kinetically facile oxidants
than the octahedral cobaltoorganic complexes, such
Ira A. Weinstock was born in Ithaca, NY, in 1959. He received a B.A.,
as cobalt(III) acetate, typically used in the oxidation
magna cum laude, from Williams College in 1988 and earned an M.A. of organic substrates. The possible effects of Jahn-
under the direction of Carlo Floriani in 1986 from Columbia University. Teller distortion on the inner-sphere reorganization
As an NSF Graduate Fellow at M.I.T., he earned a Ph.D. under the energy associated with the l-e– reduction of CIIIW125–
guidance of Richard R. Schrock in 1990. At M.I.T. he synthesized were not considered. These, and related structural
complexes containing metal-carbon, -nitrogen, and -oxygen multiple and electronic issues, will presently assume greater
bonds and studied the mechanism of alkyne metathesis by rhenium(VII) importance in evaluation of the role of inner-sphere
alkylidyne complexes. He subsequently spent one year at Sandia National
Laboratory, Albuquerque, NM, where he developed metal oxide thin films reorganization energies, and of the potential for
for applications in electronics and catalysis. In 1991 he joined the staff nonadiabatic behavior, in electron self-exchange be-
of the USDA Forest Service, Forest Products Laboratory, Madison, WI, tween CoIIIW125– and CoIIW126– and between other
the National Laboratory of the U.S. Forest Service. There, he initiated a POM anion pairs.
program in the use of water-soluble polyoxometalate salts as selective Since publication of Chester’s work in 1970,
CoIIIW125– and related POM anions have been used
catalysts for the oxidative delignification (bleaching) of commercial wood
pulp. The goal was to replace the chlorine compounds used in
papermaking by dioxygen, the least expensive and most environmentally as well-defined outer-sphere electron-transfer agents
attractive oxidant available. The program, now supported by a partnership in over 50 published studies. Indeed, the CoIIIW125–
between the U.S. Forest Service, the U.S. Department of Energy, and anion has, with good reason, been referred to as a
private industry, and which includes collaborations with the University of “soluble anode”.8 Meanwhile, the potential utility of
Wisconsin and Emory University, was recognized in 1996 with the U.S. numerous POM anions as electron-transfer agents
Secretary of Agriculture’s National Honor Award for Excellence in
Research. In addition to the design of versatile water-soluble catalyst in the selective catalytic oxidations of inorganic and
systems for the development of environmentalty sustainable chemical organic substrates of practical importance has been
technologies, the authors research interests include the kinetics and recognized. Despite these developments, the handful
themodynamics of polyoxometalate duster formation and the mechanisms of studies specifically concerned with defining the
of catalytic oxidations by dioxygen. nature of electron-transfer reactions of the POM
charge densities considerably lower than those ob- anions themselves reach conflicting conclusions re-
served in traditional low-charge-density (noncoordi- garding issues of fundamental significance.
nating) anions, such as ClO4- and NO3-; (2) the After Marcus and Sutin’s early treatments of
partial negative charges on the oxide anion ligands electron-transfer reactions, they and others continued
within POM structures are quite low (the pKa values to improve upon a number of initial approximations
of many POM anions are below 1, and the partial and assumptions in the theory.1,9–14 More than three
negative charges residing on the outermost terminal decades of interaction between theoreticians and
oxide ligands, M=O, are generally lower than those experimentalists have resulted in a dramatic refine-
on the bridging oxide anions embedded within the ment of classical and semiclassical approaches, and
clusters); and (3) many POM anions, and in particu- empirical observations have led to important conclu-
lar those possessing the Keggin structure, are close sions regarding the scope and limitations of the
to spherical in shape, a geometric consideration theoretical models.11,15–22
important for application of the electrostatic laws In parallel with progress in the theoretical arena,
upon which much of Marcus’s development was new, more accurate and versatile spectroscopic tech-
predicated. niques for observing the rates of electron self-
The first detailed kinetic study and theoretical exchange between POM anions were developed.
analysis of electron transfer involving POM anions These techniques have made it possible to obtain, and
was reported by Rasmussen and Brubaker in 1964, more thoroughly verify, 23–26 the accuracy of key
only shortly after the foundations of Marcus’s work theoretical parameters. The new kinetic techniques
had been laid.5 Rasmussen and Brubaker studied were a natural outgrowth of spectroscopic studies
the self-exchange reaction between [CoIIW12O40]6– concerning the nature of reduced POM anions (het-
(abbreviated CoIIW126–) and its l-e––oxidized deriva- eropoly “blues”).27–34 In particular, however, the use
tive, [COIIIW12O40]5– (CoIIIW125–). In addition to the of NMR to determine key theoretical parameters
general considerations listed above, both these com- followed the observation by Kozik and Baker that
plexes are substitutionally inert with respect to the l-e–-reduced heteropoly blues, although paramag-
ligand environments of the cobalt ions.6 In 1970 netic, gave sharp NMR signals.23–26 (Detailed studies
Chester recognized the potential utility of CoIIIW125– concerning the nature of reduced POM anions, al-
Electron-Transfer Reactions of POMs Chemical Reviews, 1998, Vol. 98, No. 1 115

though pertinent to electron-transfer reactions of of reduced POM anions with halogenated alkanes are
POM anions, are outside the scope of the present presented in section VI, and with inorganic oxidants
review. Nevertheless, important findings from that in section VII. Because of its importance and com-
body of work will be included as necessary.) plexity, the reduction of dioxygen by reduced POM
Given these developments, critical evaluation of the anions is treated separately in section VIII. To aid
literature concerning the nature of electron-transfer the reader, all chemical potentials have been refer-
reactions of POM anions (section III) is now possible. enced to the normal hydrogen electrode (NHE) and
Indeed, it is a necessary prerequisite to understand- all energies have been reported in units of kilocalories
ing the use of POM anions as electron-transfer probes per mole (kcal mol–1).
(i.e., “soluble anodes”) in physical or physical-organic
chemistry or as catalysts for selective oxidations. II. Outer-Sphere Electron-Transfer Reactions
First, however, some introductory comments regard-
ing theories of electron transfer are in order. AI- A. Debye-Smoluchowski and Marcus
though the interested reader will find considerably In 1942 Debye extended Smoluchowski’s method
more detail and depth in the original articles cited, for evaluating fundamental frequency factors, which
it is hoped that familiarity with the mathematical pertained to collision rates between neutral particles
forms of the theories and with the physical phenom- D and A, freely diffusing in solution, to include the
ena represented by specific mathematical relation- electrostatic effects of charged reacting species in
ships, as conveyed in the following brief introduction dielectric media containing dissolved electrolytes.35–37
(section II), will make the intellectual matter covered Debye’s colliding-sphere model assumed that the
in the present review more readily accessible to the reactant pairs formed during collisions between Dn
nonexpert. (electron donor with a charge of n) and Am (electron
Within the context of electron-transfer reactions, acceptor of charge m) possessed no inherent stability
a number of issues pertinent to analysis of reaction such that contact between colliding Dn –A m pairs was
rates of charged species in solution, such as ion- shortlived.
pairing and donor-acceptor complex formation, are When Marcus3 and others4 extended this model to
discussed. Here too, the intent is to provide the include the occurrence of adiabatic electron transfer
reader with an appreciation for phenomena com- during collisions between donor and acceptor species,
monly encountered in homogeneous-phase electron- Debye’s transient Dn – A m pairs were referred to as
transfer reactions of POM anions. “precursor” complexes ([Dn, Am ] in Scheme 1). The
Beginning with section IV, published data concern- Debye-Smoluchowski model for the diffusion-con-
ing the mechanisms of electron-transfer reactions of trolled collision frequency between Dn and Am (and
POM anions are critically evaluated. Because the between D(n +1) and A(m -1)) was retained.
primary focus is on reaction mechanisms, many Scheme 1
reports concerning stoichiometric and catalytic ho-
mogeneous-phase reactions of POM anions are not
discussed. The review is limited, rather, to discus-
sion of kinetic or other published data pertinent to
understanding the mechanisms of electron-transfer
reactions.
Because of this focus, relatively short portions of
much longer published studies are sometimes given Marcus’s theoretical treatment of outer-sphere
disproportionate attention. Nonetheless, consider- electron-transfer reactions3 related the free energy
able effort has been made to place mechanistic data of activation for electron transfer, ∆G ‡ (and thus
within the context defined by the larger significance implicitly kfet, k = rate constant, fet = forward
of the reactions under discussion. On the other hand, electron transfer), to the corrected free energy of the
POM anions have frequently been used to address reaction, ∆G°', via a quadratic equation (eq l), 4,38
issues of fundamental significance, or of considerable
interest with respect to important industrial or bio- (1)
logical processes. In some of these cases, the chem-
istry of the POM anions has taken a back seat to dis-
cussion of the conclusions reached through their use. where χ is the reciprocal Debye radius,15,37
Finally, because discussion of reaction mechanisms
requires familiarity with the behaviors of all the
species involved, brief summaries or clarifications of (2)
the chemistries of key substrates have been included.
A good example is the brief review of experimental in which D is the dielectric constant of the medium,
data and attendant theoretical considerations perti- k is the Boltzmann constant, and Σnizi2 = 2µ, where
nent to determination of the outer-sphere reorgani- µ is the total ionic strength of an electrolyte solution
zation energy of the O2/O2 •– reactant pair presented containing molar concentrations, ni, of species i of
in section VIII.C. charge z (µ ≡ (½)Σnizi2).
Oxidations of organic electron donors by POM The first term in eq 1 was retained from the
anions are covered in section IV, followed by oxida- colliding-sphere model: The electron-donor and elec-
tions of inorganic substrates in section V. Reactions tron-acceptor molecules were viewed as spheres of
116 Chemical Reviews, 1998, Vol. 98, No. 1 Weinstock

radii r1 and r2, which possessed charges of z1 and z2 , (E°), the Marcus equation (eq 1) can be fitted to a
respectively (e is the charge of an electron). This plot of ∆G ‡ vs ∆G°' by nonlinear regression, using λ
term defines the electrostatic energy (Coulombic as an adjustable parameter. A good fit between
work) associated with bringing the two spheres, ini- calculated and experimental curves for a set of like
tially infinitely removed from one another, to a dis- reactions can be used to provide evidence for an
tance r12 = r1 + r2 (formation of the precursor com- outer-sphere electron-transfer mechanism and to

plex, [Dn, Am ] in Scheme 1). The magnitude of the estimate λ for the reactions. Often ∆G is not known,
Coulombic term is modified by a factor exp(–χr 12 ), but bimolecular rate constants, k, for the electron-
which encompasses the attenuating effects of the transfer reaction have been measured. In these
dielectric medium (of dielectric constant D) and total cases, the Eyring equation,
ionic strength, µ, of the solution (eq 2).
The thermodynamic driving force for the electron-
transfer reaction was represented by the corrected
standard free energy, ∆G°', which is the standard free can be applied, where ∆G ‡ is defined by eq 1 and Z
energy, ∆G°, adjusted to account for the free-energy is the collision frequency in units of per molar per
change associated with the changes in charge, z, of second (M–1 s–1. Expansion of eq 1 gives
reactants and products, upon electron transfer be-
tween the donor and acceptor molecules in the
reaction medium:

where the Coulombic work term is abbreviated as


W(r). Substituting eq 5 into eq 4, taking the natural
From the standard reduction potentials, E°, of the logarithm, and rearranging gives
donor and acceptor species, where E° ≡ – ∆G°/nF (n
is the number of electrons transferred, and F is the
Faraday constant), eq 3 can be used to calculate ∆G°'
from electrochemical data.
The term λ (eq 1) is the reorganization energy
associated with electron transfer within the precursor For a series of reactions possessing similar λ values,
complex. Electron transfer induces changes in bond eq 6 can be fitted to a plot of ln k vs ∆ G°' by nonlinear
lengths and bond angles (in-plane and torsional) of regression, using λ as an adjustable parameter.
the donor and acceptor molecules and in the arrange- In more limited cases, the Marcus equation can be
ment of solvent molecules around the reacting pair. approximated by a linear free-energy relationship. If
These inner-sphere (bond-length and -angle) and |∆G°'| ^ λ the last term in eq 5 can be ignored, such
outer-sphere (solvent) reorganization processes are that
associated with enthalpic and entropic “costs”, which
present an activation barrier to electron transfer.
According to the Franck-Condon principle, electron
transfer within the transition-state complex occurs
at a much greater rate than bond or solvent reorga-
nization. Thus, the lowest energy pathway for elec- If the reorganization energies, λ, of a set of like
tron transfer involves a transition-state complex in reactions are similar and W(r) is small or constant,
which inner- and outer-sphere rearrangements have a plot of ∆G ‡ vs ∆G°' will be linear and have a slope
occurred so as to minimize the combined potential of 0.5. By using eq 4 and the definition of the
energies of the incipient reduced and oxidized prod- equilibrium constant, K = A exp(–∆ G°'/RT), where
ucts. The energetic cost incurred en route to the A is a constant, a plot of ln k vs ln K will be linear
transition state is the sum of-the inner- and outer- and have a slope of 0.5:
sphere reorganization energies, λ = λin + λout, of the
donor-acceptor-electrolyte-solvent system.
In contrast to the substantial bond and solvent
reorganization associated with electron transfer, very
little electronic coupling between donor and acceptor Alternatively, given that ∆G° = –nFE°, a plot of
species ( <~1 kcal mol–1) occurs along the reaction In k vs E° for the series of reactions will have a slope
coordinate leading to the transition state. This is the of 0.5(nF/RT), and a plot of log k vs E° will have a
fundamental difference between outer-sphere pro- slope of 0.5(nF)/2.303 RT or 8.5 V–1 for n = 1 at 25
cesses and inner-sphere reactions, which often, but °C. Observation of a linear free-energy relationship
not always, involve ligand exchange or atom (e.g., (cf., ln k vs ln K) whose slope is 0.5, or in the
oxygen, hydrogen, or hydride) transfer.39–44 electrochemical case flog k vs E°), 8.5 V–1, is an
indication that the electron-transfer step is outer-
B. Marcus Equation sphere in nature.3 Because of confusion in the
literature regarding the application of Marcus theory
1. Applications to intramolecular electron transfer within stable
For a series of reactions possessing similar λ values donor-acceptor complexes, alternative definitions of
and differing only in Coulombic terms and in ∆G° E° and the effect of sign convention on the slope (+
Electron-Transfer Reactions of POMs Chemical Reviews, 1998, Vol. 98, No. 1 117

or –), and terminology, these linear free-energy description of phenomena encountered in classical
relationships are discussed in greater detail in section Marcus theory.
II.D. In at least one case,23 rate constants of electron-
transfer reactions involving POM anions have been
2. Theoretical Calculation of Outer-Sphere estimated using Sutin’s semiclassical method. How-
Electron-Transfer Rate Constants ever, there is a more compelling reason for briefly
The free energy of activation (∆ G ‡) of the elemen- describing Sutin’s approach in this introductory
tary electron-transfer step of an outer-sphere reaction overview. As discussed more fully later in this
can be calculated using the Marcus equation (eq 1). section (section II.D), it must be recognized that the
The calculation of ∆G ‡ (henceforth, ∆G ‡calc) requires “equilibrium constants” described here are radial
thermodynamic data (∆ G° or information from which distribution functions and not the formation con-
(∆G°) can be calculated, such as E°, or the equilibrium stants typically associated with stable, if not always
constant, K); the dielectric constant of the solvent, readily observable, donor-acceptor reactant pairs.
D; the ionic strength, µ, of the reaction solution; the
charges, z; and radii, r, of the reactants; and a means 3. Outer-Sphere Electron Self-Exchange
for calculating λ. If the preexponential factor, 2, in The reorganization energy, 1, associated with elec-
eq 4, can be estimated or calculated, ∆G ‡calc from eq tron transfer to, or from, a molecule can be deter-
1 can be used to obtain a theoretical bimolecular rate mined experimentally from the rate constant for
constant, kcalc. Theoretical kcalc values are often electron self-exchange between that molecule and its
compared to experimentally determined rate con- l-e--oxidized or -reduced form. In self-exchange
stants of suspected outer-sphere electron-transfer reactions, such as that between Am and Am–1, ∆G°'
reactions. Good agreement between kcalc and k obs is = 0:
often viewed as evidence of an outer-sphere mecha-
nism. Poor agreement is often rationalized on the
basis of uncertainties in the assumptions required
for the theoretical calculation or is viewed as evidence In these cases, the Marcus equation (eq 1) reduces
that the electron-transfer reaction is not outer-sphere to
in nature.
Rate constants can also be calculated by means of
Sutin’s semiclassical (partly quantum-mechanical)
approaches. Here, electron transfer is handled within
a broad theoretical framework common to processes Given the charges and radii of the reactions, the
that occur rapidly on the nuclear time scale, such as dielectric constant of the solvent, and the ionic
electronic energy transfer and a number of nonra- strength of the solution, the work term, W(r), can be
diative processes. 14,45,46 Within these frameworks, estimated. Given ∆G‡ , the reorganization energy for
observed rate constants are expressed as a combina- electron self-exchange, λ11, can then be calculated
tion of diffusion-controlled rate constants for the using eq 14.
formation of transient “precursor complexes” and Alternatively, λ11 can be calculated from the ob-
activation-controlled rate constants, kact, for the bi- served rate constant, k, of the reaction shown in eq
molecular electron-transfer reaction: 13 by using eq 15, obtained by substituting eq 14 into
eq 4. For reactions in solution, Z is often taken to
be ~6 × 1011 M–1 s–1.38

The bimolecular rate constant, kact, in turn is ex- 4. The Marcus Cross-Relation
pressed as the product (eq 10) of an ‘equilibrium
constant”, K(r) (in M–1), for formation of reactant The rate constant associated with a particular self-
pairs separated by a distance r (more exactly, a radial exchange reaction (and hence the reorganization
distribution function is used, eq 11) and a first-order energy, λ22, of that self-exchange reaction) can often
rate constant, ke l (r) (in s–1, for electron transfer be inferred from the rate constant of the cross-
within the reactant pairs: reaction, k12, between the species of interest, B, and
another donor or acceptor, A, whose reorganization
energy, λ11, is known:

Provided the reorganization energy associated with


Finally, the first-order rate constant for electron the cross-reaction (λ12) is the mean of those associated
transfer is expressed as a product of nuclear, elec- with the Am /Am +1 and Bn – 1/Bn self-exchange reactions
tronic, and frequency factors: then14,45,46

The product in eq 12 is a quantum-mechanical where


118 Chemical Reviews, 1998, Vol .98, No. 1 Weinstock

largely governed by the balance of energetic factors


associated with the cations.
In analyzing the Fe(CN)63–/4– self-exchange reac-
tion, Shporer17 considered the possible effects of ion
pairing on electrostatic terms, on reorganization
energies, and on the mechanics of electron transfer
and itself. Three issues were addressed: (1) the effect of
ion pairing on the Coulombic terms associated with
bringing together similarly charged species (Coulom-
bic term in eq 1); (2) the effect of ion pairing on the
As above (eq 4), Z is the preexponential factor, and reorganization energies, λ, associated with electron
Wii are the Coulombic work terms (eq 1) associated transfer; and (3) the possibility that electron transfer
with all four combinations of the reacting species. If might occur via the ion-paired cation, with the cation
kii is known, then experimentally determined values serving as a lower energy pathway between the donor
of k12 can be used to calculate k22 and its associated and acceptor species. Of these, the first is undoubt-
reorganization energy, λ2 2 . edly important, while the second is more difficult to
assess.19 With regard to the third proposal, recent
work by Kirby and Baker suggests that simple alkali,
C. Ion Pairing between Electrolyte Ions and alkaline-earth, or tetraalkylammonium cations do not
Electron Donors or Acceptors serve as conducting bridges between POM anions.50
Catalytic and inhibitory effects of ion pairing on Kirby and Baker investigated rates of electron
the rates of outer-sphere electron-transfer reactions transfer through cationic bridges by observing the 31P
are well documented.17–19 In a study of the kinetics NMR spectra of reduced tilacunary Keggin and
of electron self-exchange between the ferri- and Wells-Dawson derivatives (heteropoly blues) rigidly
ferrocyanide anions, Fe(CN)63– and Fe(CN)64–, Shpor- linked by diamagnetic (Zn(II) or Cd(II)) or paramag-
er observed rate increases in aqueous solution cor- netic (Co(II) or Ni(II)) cations and of reduced mono-
responding to the series H+ to Cs+ and Mg2+ to Sr2+.17 lacunary Wells-Dawson anions flexibly attached via
In alkaline aqueous solutions at constant ionic a single Th(IV) atom. When the paramagnetic link-
strength, Wahl19 observed decreasing rates of self- ing cations, Co(II) or Ni(II), were present, substantial
exchange between Fe(CN)63– and Fe(CN)64– with antiferromagnetic interaction between the “blue”
increasing size of the tetraalkylammonium cations, electrons and the unpaired electrons of the paramag-
Me 4N+, Et 4N+, n-Pr4N+, n-Bu4N+, and n-Pent4N+. netic cations was observed (see original article50 for
Similar effects were observed in acetic acid.18 details of the spectroscopic evidence).
Shporer17 attributed the specific cation effects, In similarly rigid structures linked by diamagnetic
which paralleled increasing ionic radii of the elec- Zn(II) or Cd(II), no electronic interaction between the
trolyte cations, to ion pairing with the negatively reduced tungstophosphate clusters could be discerned
charged iron complexes. The factors that dictate the by 31P NMR. However, in the flexible Th(IV)-linked
extent of interaction of electrolyte cations with POM complex, syn- [(α2P2W17O61)ThIV]16–, in which the two
anions in aqueous solution were summarized nicely reduced lacunary α-[P2W17O61]10– units were in close
by Kirby and Baker:47 The ionic radii of Li+, Na+, K+, contact with one another, significant electronic in-
Rb+, and Cs+ are, respectively, 0.60, 0.95, 1.33, 1.48, teractions between electrons located on opposite sides
and 1.69 Å. However, because the smaller cations of the Th(IV) ion were observed. This interaction,
attach water molecules more firmly, the hydrated however, appeared to result from direct exchange
radii of these ions in water decrease with increasing between the reduced clusters rather than from
cation size: 3.40, 2.76, 2.32, 2.28, and 2.28 Å for the electron transfer via the Th(IV) bridge: When the
above series.48,49 Corresponding hydration numbers complex assumed the anti conformation, wherein
for Li+ through Cs+ are in the ratios: 25.3:16.6:10.5: close contact between the reduced tungstophosphate
...9.9, which reflect hydration energies in the ratios: structures was not possible, no electronic interactions
123:97:77:70:63. These data reflect larger ion-induced- were observed.
dipole terms in the interactions between the smaller Kirby and Baker’s studies do, however, point to a
cations and the negatively charged and highly po- role for electrolyte cations not mentioned by Shporer.
larizable oxygens of water. In ion-pair formation, The cations might catalyze electron transfer between
Coulombic driving forces are counterbalanced by negatively charged species by facilitating the forma-
losses in hydration energy of the cation and of the tion of ternary [Xm–-M+-X n– ] (m + n –1)– complexes. Such
POM anions. For the cations, losses in hydration complexes are particularly relevant to the study of
energies decrease with ion size, and formation con- electron-transfer reactions of highly negatively charged
stants for ion pairing increase in parallel. As a POM anions. In addition, the formation of ternary
whole, however, the POM anions possess very low complexes is particularly likely when reactions be-
charge densities and low negative charge densities tween POM anions, or between cations and other
on their surface (M=O) oxygens (M = V(V), Mo(VI), charged species, are carried out in solvents of low
W(VI); the terminal oxide, O2–, anions actually dielectric constant26,51 The application of Marcus
donate more than four electrons to the oxophilic d0 theory to outer-sphere electron transfer within bi-
addendum atoms). Thus, the extent of ion-pairing nary, or higher order, donor-acceptor complexes is
between electrolyte cations and POM anions is discussed in section II.D, immediately below.
Electron Transfer Reactions of POMs Chemical Reviews, 1999, Vol. 99, No. 1 119

Finally, ion pairing between cations (H+, alkali- ionic strength, (2) ion pairing between [CoIIIW12O40]5–
metal, alkaline-earth, or tetraalkylammonium cat- and any cations present in solution, and (3) the
ions) and POM anions can strongly influence the formation of stable complexes between [CoIIIW12O40]5–
electrochemical reduction potential, E°, of the latter.52 and electron donors. In the simplest case (1), in-
Because the rate of outer-sphere electron transfer is creasing ionic strength simply attenuates the Cou-
exponentially dependent upon thermodynamic driv- lombic term in eq 1. In the second case, [CoIIIW12O40]5–
ing force (i.e., ∆G°), small changes in reduction reacts with cations, M+ (most often, alkali-metal
potential (recall that E° = –∆G°/nF) can have large cations such as Li+, Na+, and K+), to give ion-paired
effects on reaction rates. anions of the form [(M+)(CoIIIW12O40]5–)]4–. Ion pair-
ing of this type has two consequences: It reduces the
Coulombic work term in eq 1 by decreasing the
D. Outer-Sphere Electron Transfer within Stable charge product, z1 z 2, of the reacting species, and it
Donor-Acceptor Complexes changes the reduction potential, E°, of the POM
anion, thus altering the thermodynamic driving force,
The Marcus equation (eq 1) and its derivatives (eqs ∆G°'.
7, 14, and 17), as well as Sutin’s semiclassical Both ionic strength and ion pairing can impact the
approach, were based upon a colliding-sphere model third phenomenon: the formation of stable complexes
in which reactant pairs (precursor complexes) pos- between [CoIIIW12O40]5– and electron donors. Evalu-
sessed no inherent stability. Sutin has noted, how- ation of the effects of the formation of donor-acceptor
ever, that although preequilibrium formation of complexes on the relationship between experimen-
precursor complexes was not explicitly postulated in tally measured kobs values and bimolecular rate
Marcus’s formalism, “stability constant” of precursor constants, ket, predicted by the Marcus equation
“complexes” were implicated by his mathematical requires consideration of the reaction steps and
treatment.53 In Sutin’s semiclassical treatment (eq associated rate constants shown in Scheme 2.14,54–56
10), these “equilibrium constants” were described Unlike the transient ion pairs [Dn, Am ] and
more explicitly. However, despite the mathematical
[D(n +l), A(m -1)] described in Scheme 1, the ion pairs
similarity of radial probability distribution func- described in Scheme 2, [Dn •••Am ] and [D(n +1)•••A m–1 )],
tions13,46 (with units of M–1) to formation constants refer to association complexes with negative enthal-
of stable donor-acceptor pairs, no departure from the pies of formation.
colliding-sphere model is intended.
At the same time, the quadratic dependence of Scheme 2
∆G ‡et on ∆G°' and λ, developed by Marcus (and
inherent in Sutin’s quantum-mechanical treatment)
and which describes the activation barrier to outer-
sphere (adiabatic) electron transfer, is applicable
whether electron transfer occurs within transiently
formed reactant pairs (colliding-sphere model) or
within stable donor-acceptor complexes. In each of
these cases, however, kobs values measured for the
overall electron-transfer reactions have different In the most general case, no restrictions are placed
meanings. In the classical (colliding-spheres) case, on the values of rate constants or associated equi-
k obs is equivalent to ket, the bimolecular rate constant librium quotients: K1 = k1 /k–1, Ket = kf e t /k ret, or K3 =
described by the Marcus equation (eq 1), or to k act in k 3 /k–3, where kfet and kret are the rate constants for
Sutin (eq 10). For adiabatic (outer-sphere) electron forward and reverse outer-sphere electron transfer,
transfer within stable donor-acceptor complexes, the respectively. The rate constant for formation of
relationship between kobs and ket (or kact) is not always donor-acceptor pairs of formation constant K1 is
straightforward. This point has caused confusion given by k1; the rate constant for dissociation of the
among some who have used POM anions, particularly successor complex, [D(n +1)•••A( m -1)], formed via in-
[CoIIIW12O40]5–, as outer-sphere oxidants. tramolecular electron transfer within [Dn •••A m ], is k3 .
In particular, lack of agreement between kc a l c Usually, however, one of three situations, depend-
values obtained using the Marcus equation (eq 1) and ing on the energetics of the electron-transfer step,
experimentally determined rate constants, kobs, need applies:57 Case 1. The electron-transfer step is high-
not imply that electron transfer is inner sphere in ly exergonic and rapid, such that kfet > k1, kfet > k–1,
nature; intramolecular outer-sphere electron transfer and k3 p kret. Case 2. The electron-transfer step is
within stable donor-acceptor pairs must also be slightly endergonic or exergonic (|∆G°et| is small) and
considered. A brief evaluation, using [CoIIIW12O40]5– reversible (k re t ≈ kfet ^ k1 and kret ≈ k fet ^ k3). Case 3.
as an example, follows. The electron transfer step is substantially endergonic
The Co(III) ion in [CoIIIW12O40]5– is substitutionally and reversible, such that kfet ^ k1 and kret p k 3 .
inert. Thus, it is reasonable to expect that this ion Case 1. In this case, the reaction is diffusion con-
will be reduced via outer-sphere pathways. At the trolled and kobs = k1 k1[D n ][A m ]. In this regime, no corre-
same time, the [CoIIIW12O40]5– anion possesses a high lation between kobs and ∆G°fet (or E°fet) is observed.
negative charge. As a result, three phenomena that Case 2. The reaction is activation controlled: ko b s
might complicate the kinetic analysis of reactions of = K1 k fet[D n ] [Am ]. The term “activation control” refers
this and related anions with electron donors must to the dependence of kobs on kfet. Despite this depen-
be considered. These phenomena are (1) effects of dence, however, kobs does not obey the Marcus equa-
120 Chemical Reviews, 1998, Vol. 98, No. 1 Weinstock

tion (eq 1). There are two reasons why. First, the Combining eq 21 and the Eyring equation (eq 4) gives
Coulombic work term, W(r), in eq 1, taken from the
colliding-sphere model, does not apply here. Rather
than a transient ‘“precursor” complex such as that
depicted in Scheme 1, [Dn •••A m ] is a stable complex in which nonessential constants have been omitted.
whose concentration at equilibrium is dictated by K1 . The equilibrium constant, K, is defined by a similar
The remainder of eq 1, however, which includes the exponential relationship:
reorganization energy, λ, and describes the quadratic
dependence of ∆G ‡fet (and hence kfet on ∆G°fet, is still
applicable. A complication arises, however, in that The natural logarithms of eqs 22 and 23 are
k obs is not equal to kfet. Nonetheless, provided K1 is
independent of ∆G° for the electron-transfer reaction,
then kobs is linearly proportional to kfet over a range
of ∆G° values. Making use of this relationship, the and
quadratic portion of the Marcus equation can be used
to evaluate kobs (eqs 20-31).
Case 3. In this case, the reaction is under ther- Thus, returning to Scheme 2, if kobs = kfet, then
modynamic control. Because kret p k3, kobs = K1 K etk3
and kobs = Ket (i.e., kf e t /k ret). However, while directly
proportional to Ket, and hence under thermodynamic
control, kobs is also a linear function of k3. For this
reason reactions of this type, while very different such that a plot of ln kobs vs ln K gives a straight line
from those described by case 1, are sometimes also with a positive slope of 0.5.
referred to as “diffusion controlled”.57 When encoun- Because electrochemical reduction potentials are
tered in the literature, reference to reactions of the often experimentally accessible, it is useful to con-
type described in case 3 as “diffusion controlled” can sider the above relationships in electrochemical
lead to confusion. terms. The standard thermodynamic potential of a
In cases 2 and 3, in which kobs is proportional to reversible galvanic cell (electromotive force, emf) or
kfet, determination of kfet requires knowledge of K1 of an electron-transfer reaction (i.e., the sum of two
(case 2) or of K1, Ket, and k3 (or more generally K3 ) half-reactions) is defined as
(case 3). For reactions between charged species, K1
and k3 or K 3 (each functions of the parameters
associated with the Coulombic work terms defined and
in eqs 1 and 2) can be evaluated both theoretically58
and experimentally. Experimentally, values of K1
and Ket can be obtained by fitting integrated rate
expressions to raw kinetic data by means of nonlinear Thus, returning to eq 25,
regression techniques in which K and/or k are adjust-
able parameters .59 In unique cases, such as in the
oxidative electron-transfer luminescence quenching and using eq 24 for ln k,
of photoexcited donors within donor-acceptor ion
pairs, kfet can be measured more directly.60,61
In case 1, no correlation between kinetic and
thermodynamic parameters is observed. In cases 2
and 3, however, linear relationships exist between For n = 1 and in log10
specific functions of kobs and ∆G° (or E°), and the
slopes of the lines obtained by plotting these func-
tions can be used to differentiate between the two
cases:
Case 2: kobs = kf e t . Because kobs is proportional Thus, if kobs = kfet, plot of log kobs vs E° gives a straight
to kfet, the quadratic term in the Marcus equation (eq line with slope of +8.5 V–1.
1) applies. Because the electron transfer is occurring Case 2: kobs 0~ Ket. By definition,
within a stable complex, the first term in the Marcus
equation drops out, as does the need to apply an
electrostatic correction to ∆G°. This leaves only the
quadratic dependence of ∆G ‡ on ∆G°: Because kobs is proportional to Ket, rather than to kfet ,

such that
For limited ranges of ∆G°, an expression analogous
to eq 7 applies:

Plot of log kobs vs E° now gives a straight line with a


slope of +16.9 V–1.
Electron-Transfer Reactions of POMs Chemical Reviews, 1998, Vol. 98, No. 1 121

Note that the signs of the slopes of lines obtained


in plots charted using electrochemical data can be
either positive or negative depending on how E° is
defied. The above relationships were derived in
terms of the standard thermodynamic potentials, E°,
of electron-transfer reactions. By convention E° =
54,55
–(E°donor – E°acceptor). (In the more general case
E°donor or E°acceptor can be electrode potentials).57 By
definition (i.e., eq 27), if an electron-transfer reaction
is spontaneous, E° is greater than zero. Hence, if for
a series of outer-sphere electron-transfer reactions
involving a common donor or acceptor species, one
plots log kobs vs E°, –1the slope of –1the plot will be
positive (i.e., +8.5 V or +16.9 V ).
However, values of log kobs for electron-transfer
reactions of a given species are also plotted as a Figure 1. Plot of ln k vs exp(–χr12) (eq III1) for the elec-
function of standard reduction potentials (also E°) , tronIII self-exchange
6– 5
reaction between [Co W 12O 40]5–
and
[Co W12O40] .
of individual electron-donors or acceptors (usually a
set of donors or acceptors are used). This is equiva-
lent to plotting log kobs as a function of a varying
electric constant, D, of the reaction medium was
electrode potential. In these cases, slopes of lines
reduced by addition of dioxane. In addition, at
obtained in plots of log kobs vs E° can be either
constant ionic strength, replacement of Li+ ions by
positive or negative, depending on one’s frame of
K+ ions resulted in a nearly 10-fold increase in k.
reference. For example, if the species under inves-
tigation is an electron donor, rates of electron transfer From a plot of ln k vs the Debye term, exp(–χr1 2 )
to the acceptors increase as the standard reduction (eq 1), an experimental value of k0 = k(µ=0M) = 4.5 ×
potentials (E° of the acceptors shift to more positive 10–3 M–1 s–1 was obtained by extrapolation to
values. Hence, the sign of the slope of the plot of log exp(–χr12) = 1 (µ= 0 M). From the slope of the plot
kobs vs E° is positive. However, if the species under of ln k vs exp(–χr12), a charge product of z1 z2 = ~8
investigation is an electron acceptor, rates of electron was calculated (Figure 1).
transfer increase as the E° values of the donors shift Had the reactions been carried out within the ionic
to more negative values. The sign of the slope of the strength limit of the Debye equation (µ < ~0.001 M),
plot of log kobs vs E° is now negative (i.e., -8.5 V–1 or a charge product of 30, corresponding to z1 z2 =
– 16.9 V–1). (–5)(–6), would have been expected. Nonetheless,
in calculation of a theoretical value for k(µ=0M), the
experimentally determined charge product of 8 was
III. Electron Self-Exchange between POM Anions used to calculate the Coulombic contribution to ∆G‡
at infinite dilution (W(r) in eq 1 with z1 z2 = 8 and
A. Keggin Tungstocobaltate Anions exp(–χr12) = 1), which gave W(r)(µ=0M) = 3.0 kcal
mol –1.
Rasmussen and Brubaker carried out the first The inner-sphere reorganization energy, λin = 11.05
direct measurements of self-exchange rates for elec- kcal mol–1 (converted from the published value of
tron transfer between POM anions in water using 7.68 × 10–13 erg), was calculated using Pauling
[CoIIW12O40]6– (CoIIW126–) and its one-electron-oxi- crystal radii after a ligancy correction from six to
dized derivative [CoIIIW12O40]5– (CoIIIW125–).5 The four: 1.97 Å for four-coordinate CoII–O and 1.87 Å
reduced anion, CoIIW126–, first synthesized and char- for four-coordinate COIII–O. The latter value com-
acterized by Baker and co-workers,62–64 was prepared pared well with that determined by X-ray diffraction
using small amounts of 60Co, an unstable isotope of for K5 [CoIIIO4W12O36].65 An outer-sphere reorganiza-
cobalt. The extent of electrontransfer from CoIIW126– tion energy of 18.4 kcal mol–1 was obtained using a
to CoIIIW125– was determined after selectively pre- [CoW12O40] n – radius of 5.0 Å. The total calculated
cipitating CoIIW126– by addition of tetra-n-butylam- reorganization energy was 29.5 kcal mol–1.
monium iodide. Rates of electron transfer were then These values gave ∆G ‡(µ=0M) = 10.4 kcal mol–1 (eq
calculated from the specific activities of the insoluble 1). Equation 4, with Z = 1011 M–1 s–1, gave a
tetra-n-butylammonium salts of CoIIW126–. At 0 °C, theoretically predicted value of k0calc = 4.7 × 102 M–1
pH 2 (HCl), and an ionic strength of µ = 0.6 M (LiCl), s–1, which was 105 to 106 greater than that obtained
the reaction was first-order in both CoIIW126– and experimentally (k0 = 4.5 × 10–3 M–1 s–1).
CoIIIW125– with k = 0.63 M–1 s–1 (eq 35). Rasmussen and Brubaker suggested that the dis-
parity between experimentally observed and calcu-
lated k values might have resulted from the use of
too many simplifying assumptions. Alternatively,
The rate constant was independent of [H+] over a the authors proposed that because the cobalt cations
pH range of 0.10 to 1.6 but showed a marked were embedded in [W12O40]8– shells, electron transfer
dependence on ionic strength, k(µ=0.015M) = 0.03 M–1 between the Co(II) and Co(III) ions might not have
s–1 and k(µ=0.81M) = 0.93 M–1 s–1. The rate constant met the requirements of the theory. These alterna-
for the self-exchange reaction decreased as the di- tives are further evaluated in section III.C.
122 Chemical Reviews, 1998, Vol. 98, No. 1 Weinstock

B. Keggin and Wells-Dawson Metallophosphate [P2Mo3W15O62]6– and [P2Mo3W15O ]7– corresponded


3 62 –1 –1
Anions to an exchange rate of ~2 × 10 M s , about 1
order of magitude smaller than the diffision-
The next direct measurements of the rates of controlled rate limit. The slower rate was attributed
outer-sphere electron self-exchange between POM to localization of the blue electron in the Mo3O13 triad.
anions were reported in 1986.25 Rate constants were Well-resolved 17O NMR spectra obtained from
calculated by analysis of the NMR spectra of equi- acetonitrile solutions of oxidized and l-e--reduced
molar mixtures of fully oxidized and 1-e--reduced POM anions (free-acid forms) were also used to
POM anions (1-e- heteropoly blues). On the basis estimate POM self-exchange rate constants.69 How-
of line broadening observed in 31P NMR spectra, ever, only limiting (minimum) values were ob-
rate constants were determined for self-exchange be- tained: At room temperature, a rate constant of
tween the Keggin anions α-[PW12O40]3– (PW123–) and greater than ~5 × 103 M–1 s–1 was estimated for self-
l-e--reduced α-[PW12O40]4– (PW124–), between the exchange between [PMo12O40]4– and [PMo12O40]3–,
Wells-Dawson anions α-[P2W18O62]6– (P2W186–) and and a rate constant of greater than ~4 × 104 M–1 s–1
α-[P2W18O62]7– (P2W187–), and between the trimo- was estimated for self-exchange between [SiW12O40]5–
lybdenum-capped derivatives α-[P2Mo3W15O62]6– and and [SiW12O40]4–. At 65 °C the rate constant for
α-[P2Mo3W15O62]7–. (In 1975 rateV data for4–elec- electron transfer between the metatungstates
tron self-exchange between [PMo W11O40] and [H2W12O40]7– and [H2W12O40]6– was estimated to be
[PMoVIW11O40]3– in nonaqueous solvents was ob- greater than ~4 × 104 M–1 s–1. Because the nature
tained using electron-spin resonance (ESR) spectros- of solutions of the free acids of POM anions in
copy.51 Observed rate constants were reportedly acetonitrile (extent of ion pairing, diffusion coef-
greater than those predicted by diffusion control, a ficients, etc.) were not reported, it is difficult to draw
result that suggested25 the formation of charge- conclusions from these data. However, these results
transfer complexes between the POM anions.) are consistent with those obtained by analysis of the
A key difference between self-exchange between 31
P NMR spectra of aqueous solutions of related
CoIIW126– and CoIIIW125– and between PW124– and tungstophosphates.25
PW123– is that in CoIIW126– the exchanged electron In a detailed investigation,23 Kozik and Baker
is localized at the center of the POM anion within performed a variable ionic strength study of the rates
the d orbitals of a tetrahedrally coordinated Co(III) of electron transfer between PW123– and l-e--reduced
ion. On the basis of spectroscopic (ESR) and mag- PW-124– and between PW125–4– and the 5–diamagnetic
netic properties of 1- and 2-e--reduced [CoIIW12O40]6– 2-e -reduced anion PW12 . In PW12 the two d
(i.e., [CoIIW12O40]7– and [CoIIW12O40]8–), it was con- electrons, although residing on different tungsten
cluded that the central (Co(II)) heteroatoms are atoms and thus widely separated from one another,
effectively electronically isolated from the reduced are completely paired by a “multiroute super-ex-
tungsten framework. 30 By contrast, in the 1-e-- change” mechanism. 23,24,26,31 A more limited study of
reduced tungstophosphate anion, PW124–, the ex- the rate of self-exchange between α-[P2W18O62]6– and
changed electron, located in a predominantly non- l-e--reduced a-[P2W18O62]7– at several high ionic
bonding b2 orbital of the C4v distorted WO6 octahedron, strengths was also performed.” Theoretical outer-
is delocalized over all 12 tungsten atoms (rates of sphere rate constants were predicted using Sutin’s
intramolecular electron transfer are ~1011 to 1012 s–1 semiclassical method.14,45,46
at room temperature) via a thermal hopping mech- Rate constants for 1:1 mixtures of α-[PW 12 O 40 ] 3–
anism.25,33,66 In the 1-e--reduced Wells-Dawson and α-[PW12O40]4– (1.0 mM of each in their free-acid
anion P2W187–, the blue electron is rapidly delocalized forms), the most thoroughly studied system, were
(1.7 × 1011 s–1) over only the 12 “belt” tungsten atoms determined by 31P NMR over a range of ionic strengths
of the complex,25,33 and in α-[P2Mo3W15O62]7– delo- (µ = 0.026–0.616 M). Ionic strength was adjusted
calization of the added electron is limited to the by addition of HCl and NaCl (pH values ranged from
trimolybdenum cap and is 25 times slower than in 0.98 to 1.8). Observed rate constants, kobs, were fitted
P2W187–.25 to the Debye-Hückel equation70 (eq 36), which is
In a preliminary investigation, 31P NMR spectra conceptually related to the Debye term, exp(–χr12,
were acquired from solutions prepared from the free in eq 1. An internuclear distance of r = 11.2 Å or
acids of PW123– and PW124– (pH 1), from the lithium twice the ionic radius of the PW12n – anions, was used;
salts of P2W186– and P2W187– (pH 1 and pH 4), and at 25 °C in water, α = 0.509 and β = 0.329 Å–1.
from the potassium salts of α-[P2Mo3W15O62]6– and
α-[P2Mo3W15O62]7– (pH 2.5). No ionic-strength effects
were seen when the overall concentrations of equimo-
lar mixtures of PW123– and PW124– were increased
from 0.002 to 0.02 M. On the basis of reported A linear relationship was observed (correlation
diffusion coefficients,31,67,68 the rate constant for coefficient of 0.998), from the slope of which a charge
electron transfer between the Keggin anions ap- product of z1 z2 = 12.5 (12.0 is the theoretical value)
peared to be diffusion controlled (i.e., greater than was calculated. The linearity and close-to-theoretical
~107 M–1 s–1, based on a hydrodynamic radius of 5.6 slope was surprising in that the Debye-Hückel
Å)67 as was7–that for electron transfer between P2W186– expression (eq 36) for rates of reaction in solution was
and P2W18 (greater than ~104 M–1 s–1, based on a derived for univalent ions at low ionic strengths (up
hydrodynamic radius of 10 Å).68 The line broaden- to 0.01 M). Agreement at much higher ionic strengths
ing observed in 31P NMR spectra of mixtures of (greater than 0.5 M) was attributed to the fact that
Electron-Transfer Reactions of POMs Chemical Reviews, 1998, Vol. 98, No. 1 123

POM anions,71–73 “owing to the very pronounced Keggin anions were estimated from IR data in
inward polarization of their exterior oxygen atoms, combination with normal coordinate analyses, which
have extremely low solvation energies and very low was reported for two hexametalate anions.74 With
van der Waals attractions for one another”.23 For the the use of eq 38, a value of λin = 6.1 kcal mol–1 was
α-[PW12O40]4–/ α-[PW12O40]5– system, fairly good agree- calculated. The authors suggested that if small
ment with the Debye-Hückel expression was ob- changes in bond angles had been included, a slightly
served up to µ = 0.125 M (slope consistent with the higher λin value would have been obtained and
predicted charge product of z1 z2 = 20) before signifi- slightly bettor agreement with experimentally de-
cant deviation from linearity was observed. termined rate constants achieved.
Agreement with the Debye-Hückel expression, Calculation of the outer-sphere reorganization
and slopes near theoretically predicted values, indi- energy is straightforward, provided that accurate
cated that the “precursor complexes” (contact ion values of the hydrodynamic radii of the reacting
pairs) that formed prior to electron transfer likely species are available (eq 39).
involved “free” aqueously solvated α-[PW12O40]3–,
α-[PW12O40]4–, and α-[PW12O40]5– anions rather than
alkali metal cation paired species such as those
shown in eq 37.
In eq 39, e is the charge transferred during reac-
tion; η is the refractive index, and Ds, the dielectric
constant of the reaction medium; r1 and r2 are the
radii of the reacting species; and r12 is the inter-
nuclear distance at which the probability of electron
Theoretically predicted self-exchange rate con- transfer is at a maximum (usually set equal to r1 +
stants for the α-[PW12O40]3–/ α-[PW12O40]4– system r 2). It is noteworthy, because not intuitively obvious,
were k0 = 3.6 × 105 M–1 s–1 at infinite dilution (µ = that λout is neither an explicit function of the ionic
0 M) and k(µ = 0.616M) = 1.0 × l08 M–1 s–1 at µ = 0.616 strength in a given reaction medium nor of the
M. The effect of added electrolyte (higher ionic charges of the reacting species. Thus, unlike the
strength) was to decrease electrostatic repulsion Coulombic work term, which is highly sensitive to
between the negatively charged POM anions, which charge- and ionic-strength effects, the contribution
attenuated the contribution of the Coulombic work of outer-sphere reorganization energy to the free
term to ∆G ‡ for electron transfer (eq 1). Experimen- energy of activation, ∆G ‡, is often, to first approxima-
tally determined values were k0 = 1.1 × 105 M–1 s–1 tion, regarded as a constant property of the reactant
(by7 extrapolation using eq 36), and k(µ = 0.616M) = 8.2 × pair. For the α-[PW12O40]3–/ α-[PW12O40]4– system at
10 M–1 s–1. 25 °C in water, a value of λ out = 16.4 kcal mol–1 and
It has been stated that “the crux of the electron- a total reorganization energy of λ = λin + λout = 22.3
transfer problem is the fact that the equilibrium kcal mol–1 were calculated.
nuclear configurations of species changes when it Classically, the bimolecular rate constant of an
gains or loses an electron”.46 It is not surprising, outer-sphere electron-transfer reaction is sometimes
then, that inconsistent conclusions reached concern- written as shown in eq 40, where Z is again the
ing the outer-sphere electron-transfer reactions of universal frequency factor, and K is a factor that is
POM anions can, to a large extent, be attributed to unity for adiabatic electron transfers and less than
assumptions made with regard to pertinent inner- unity for nonadiabatic ones:13
sphere reorganization energies. The physical proper-
ties of the oxidized and reduced anions α-[PW12O40]3–
and α-[PW12O40]4– have been well studied. Thus,
Kozik and Baker’s detailed analysis of the inner- The commonly used value of Z (usually taken to
sphere reorganization energy associated with this be ~1011 M–1 s–1 in solution) is based on a colliding-
system23 provides a useful starting point for analysis sphere model in which reaction does not occur until
of other, more complex, heteropolyanions. the reactants “touch” one another (i.e., reach an
Reorganization energies for the α-[PW12O40]3– /α- internuclear distance of r12 = r1 + r2). However,
[PW12O40]4– system were calculated according to adiabatic electron transfer can occur at internuclear
literature methods. 53 The inner-sphere reorganiza- distances greater than r. Using a radial distribution
tion energy, λin, is a function of bond-length and function to modify Z accordingly, Marcus showed that
-angle changes and their associated force constants for exothermic adiabatic reactions between reactants
(eq 38, d1 and d2 are atomic coordinates, and f is a with radii of 3 Å, a value of Z = ~1012 might be more
force constant). appropriate.13 Larger Z values were predicted for
exothermic adiabatic electron transfer between re-
actants with larger radii.
A classical approach to the calculation of the rate
Bond lengths of oxidized and reduced Keggin anions constant at infinite dilution, k0, for the α-[PW12O40]3–/
were obtained from X-ray crystallographic data. On α-[PW12O40]4– system (i.e., eq 1, using Z = 1011, λ =
the assumption that λin would result primarily from 22.3 kcal mol–1, and r = 11.2 Å gives k0calc = 3.5 ×
changes in W–O bond lengths, slight changes in bond 103 M–1 s–1, substantially lower than that obtained
angles upon reduction were ignored. Force constants by Kozik and Baker using the most current level of
of individual W-O bonds in oxidized and reduced theory (k0calc = 3.6 × 105 M–1 s–1).23 Although derived
124 Chemical Reviews, 1998, Vol. 98, No. 1 Weinstock

for exothermic reactions, modified classical treat-


ment,13 using a theoretically redicted value of Z =
3.5 × 1012 M–1 s–1 (r = 11.2 Å), gives k0calc = 1.2 ×
105 M–1 s–1, which is within experimental error of
the observed value of (1.1 ± 0.2) × 105 M–1 s–1. For the α-[P2W18O62]6–/ α-[P2W18O62]7– system, ob-
The excellent agreement between theory and ex- served rate constants, k(µ=0.59M) = 1.1 × 103 M–1 s–1
periment lends support to the validity of the assump- and k(µ=1.49M) 1.4 × 104 M–1 s–1, were approximately
tions made by Kozik and Baker. In particular, in 3 orders of magnitude smaller than those predicted,
combination with the unexpectedly high degree of fit k (µ=0.59M) = 9.9 × 106 M–1 s–1 and k(µ=1.49M) = 5.5 ×
of their data to the Debye-Hückel equation at higher 107 M–1 s–1. (Rates at lower ionic strengths were too
than anticipated ionic strengths, the observed agree- slow to be determined from 31P NMR line widths.)
ment between theory and experiment suggests that The small rate constants could not be attributed
their assignment of λ = 22.3 kcal mol–1 is correct or to ion pairing (eq 37). By effectively decreasing the
very close. In addition, their application of semiclas- charge product, z1 z 2, of the reacting pair (eq 1), ion
sical theory of outer-sphere electron transfer sets a pairing would have exponentially increased observed
standard for future work in this area. The large reaction rates (eq 42, (z1 z2)ip < z1 z2):
difference between the classically calculated value (Z
= 1011 M–1 s–1) and the results obtained using
mathematical models that allow for electron transfer Rather, the small rate constants were attributed
over distances greater than the sum of ionic radii to nonuniformity in the distribution of electron
further suggests the possibility that conclusions density over the reduced Wells-Dawson anion,
reached using less-comprehensive treatments might α-[P2W18O62]7–. Not only is the anion ellipsoidal
be inadequate. rather than spherical, but the blue electron is delo-
An additional conclusion reached was that “elec- calized only over the 12 belt tungsten atoms at the
tron delocalization of itself does not significantly midsection of the anion. Both factors might be
contribute to the activation energy” of electron trans- expected to add a geometrical constraint on orienta-
fer.23 This phenomenon is not without precedent. tion of the anions during electron transfer. Although
The rate of self-exchange between [Ru(bpy)3]2+ and not evaluated further from a theoretical point of
[Ru(bpy)3]3+ (r = 13.6 Å, compared to 11.2 Å for view, this hypothesis was supported by previously
exchange between Keggin anions) is consistent with obtained experimental data. The rate of self-ex-
an adiabatic (i.e., outer-sphere) electron-transfer change between the potentially more “orienta-
mechanism. However, based on direct 4d-4d overlap tionally challenged” anions α-[P2Mo3W15P62]6– and
between the two metal centers, ab initio calculations α-[P2Mo3W15P62]7–, in which the blue electron is
showed that the reaction should be highly nonadia- localized in the trimolybdenum cap, was ~25 times
batic.53 On this basis it was concluded that electron slower than that in the α-[P2W18O62]6–/ α-[P2W18O62]7–
exchange between the metalloorganic cations oc- system. 25,26
curred predominantly via ligand-ligand overlap. The In view of these arguments, the rate of electron
observed adiabaticity was the result of π*–π* inter- self-exchange between Ru(III) and Ru(II) in the Ru-
actions between the two reactants, which arose from substituted tungstophosphate anions [PRuIII(OH2)
delocalization of the metal dπ electron density onto W11O39]4– and [PRuII(OH2)W11O39]5– was one of sev-
the π∗ orbitals of the bipyridine ligands.53 eral lines of evidence used by Rong and Pope to argue
Theoretically predicted rate constants for exchange that n-electron density from the d orbitals of the Ru-
between the l- and 2-e--reduced anions α-[PW12O40]4– (II) (d6) ion was partially delocalized onto the poly-
α-[PW12O40]5– were k0 = 2.2 × 103 M–1 s–1 and tungstate ligand. 75 A self-exchanging system was
prepared by dissolving cesium salts of the anions (1.3
k(µ= 0.125M) = 3.1 × 106 M–1 s–1. Experimentally and 0.06-0.26 mM for the Ru(III) and Ru(II) anions,
determined values were k0 = (1.6 ± 0.3) × 102 M–1
respectively) in water at high ionic strength (µ = 1.5
s–1 (by extrapolation) and k(µ= 0.125M) = 2.5 × 105 M–1
M, Na2SO4 buffer) at pH 3. A preliminary value of
s–1 (the apparent ionic-strength limit for reasonable the rate constant for electron self-exchange, k(µ = 1.5M)
fit to the Debye-Hückel equation). = 1.2 × 106 M–1 s–1, was determined by analysis of
The smaller-than-predicted experimental rate con- line broadening in 31P NMR spectra of the mixed
stants were attributed to the additional energy POM system. This value is close to that measured
needed to decouple the paired electrons in diamag- by Kozik and Baker for the similarly charged 1- and
netic 2-e--reduced α-[PW12O40]5–. The Coulombic 2-e--reduced anions α-[PW12O40]4–/α-[PW12O40]5– where
repulsion commonly associated with the formation of k(µ = 1.025M) = 2.5 × 106 M–1 s–1.
localized electron pairs is essentially absent in this Although a provocative preliminary result, the
system and not expected to contribute to the free value of this comparison is limited by (1) the high
energy of activation. The magnitude of the energy ionic strength used in the Ru system (Kozik and
likely required to decouple the paired electrons in Baker observed deviation from Debye-Hückel ionic-
α-[PW12O40]5– was not evaluated. However, from the strength dependence at µ values near 1.0 M);23 (2)
difference between experimental and theoretical the apparent use of Na+ cations in the sulfate buffer
values of k0, electron coupling would appear, from eq (Kozik used Li+ cations); and (3) the use of cesium
41, to have contributed an additional 1.45-1.76 kcal salts of the Ru complexes as opposed to the free-acid
mol–1 to ∆G‡ . salts used by Kozik (thermodynamic constants for the
Electron-Transfer Reactions of POMs Chemical Reviews, 1998, Vol. 98, No. 1 125

formation of cation–POM ion pairs increase in the


order H+ < Li+ < Na+ < K+ < C8+).52

C. Reevaluation of the Reorganization Energy of


the [CoIIIW12O40]5–/[CoIIW12O40]6– Couple
Over the past several decades, Rasmussen and
Brubaker’s published data5 have been used to calcu-
late the reorganization energy for electron self-
exchange between [CoIIIW12O40]5– and [CoIIW12O40]6–.
Calculated reorganization energies vary widely, from
λ = 23 to 61 kcal mol–1.59,76-78 During this time,
[CoIIIW12O40]5– has been used repeatedly as a well-
defined probe of outer-sphere electron-transfer reac-
tions (more than 50 published studies). In those
cases in which kinetic data are compared to the
results of theoretical analyses (Marcus), use of a
correct value for the reorganization energy of the
[CoIIIW12O40]5–/[CoIIW12O40]6– system is critical.
In particular, rates of cross-reactions between of ln k vs exp(–χr 12) (Figure 1). The slope of the line
[CoIIIW12O40]5– and organic or inorganic substrates shown in Figure 2 gives a charge product of z1 z2 =
have frequently been used to “measure” the reorga- 11.7. Extrapolation to µ = 0 M gives k 0 = 2.75 ×
nization energies associated with oxidation of the 10–3 M–1 s–1, very similar to the value obtained from
substrate under investigation (eqs 43 and 44): the plot in Figure 1. Both functions, exp(–χr 12) and
the Debye-Hückel equation, were derived for solu-
tions much more dilute than those actually used.
Hence, failure to obtain expected charge-product
or values need not of itself be viewed as evidence of ion-
pair formation.
Rasmussen and Brubaker also measured the effect
of the dielectric constant of the reaction medium on
where λ12 is the experimentally determined reorga- k at a constant ionic strength of 0.6 M. The plot of
nization energy of the cross-reaction between species In k vs 1/D, where D is the dielectric constant of the
1 and 2, while λ11 and λ22 are the reorganization reaction medium, is expected to give a slope directly
energies corresponding to self-exchange reactions proportional to the charge product of the reacting
between oxidized and reduced forms of 1 and between species (eqs 45 and 46, from eq 15):
oxidized and reduced forms of 2. When [CoIIIW12O40]5–
is used as a probe to determine the reorganization
energy, λ22, of a substrate, the result is linearly pro-
portional to the value of the self-exchange reorgani-
zation energy, λ11, of the [CoIIIW12O40]5–/[CoIIW12O40]6– from which
system (eq 44). Without reliable “calibration” (i.e.,
accurate measurement of λ11, the value of the
[CoIIIW12O40]5– “probe” is compromised. Past efforts
to derive λ11 from the data published by Rasmussen
and Brubaker can now be evaluated in light of the The slope of the plot of ln k vs 1/D is equal to
detailed studies of Kozik, Hammer, and Baker. –(z1 z2e 2 /r) exp(–χr12)/RT, from which the charge prod-
Disagreement regarding the value of λ11 for the uct, z1 z2, can be calculated. Rasmussen and Brubak-
[CoIIIW12O40]5–/[CoIIW12O40]6– couple appears to stem er’s data are plotted in Figure 3.
from an anomalous dependence of the self-exchange The slope of the line obtained by linear regression
rate constant on ionic strength. The plot of ln k vs of the data plotted in Figure 3 gives a charge product
exp(–χr12), reported by Rasmussen and Brubaker, of z1 z2 = 32. This result might be viewed as empirical
was linear and thus appeared meaningful and useful evidence for a lack of significant ion pairing with Li+
(Figure 1). At the same time, the slope of this plot cations, even at relatively high ionic strength (µ =
gave a charge product of z1 z2 = 7.8. The Debye 0.60 M). Within error, the activation energy for the
correction term, exp(–χr 12) was derived for very self-exchange reaction varied with the dielectric
dilute solutions of ionic strengths less than 0.001 M. constant of the reaction medium as expected for a
At higher ionic strengths, the Debye-Hückel relation precursor complex possessing a charge product of 30.
(eq 36), which was derived for ionic strengths of up This result may not necessarily contradict that
to 0.01 M, often gives better results. In Figure 2, obtained from the plot of ln k vs exp(–χr12) (Figure
Rasmussen and Brubaker’s data are plotted using the 1). Careful inspection reveals that the ionic strengths
Debye-Hückel equation. reported by Rasmussen and Brubaker were calcu-
Linear regression analysis of these data (dotted lated incorrectly: Some of the ionic-strength values
line) gives a coefficient of correlation almost identical reported are smaller than those dictated by the
with that obtained in Rasmussen and Brubaker’s plot concentrations of [CoIIIW12O40]5– and [CoIIW12O40]6–
126 Chemical Reviews, 1998, Vol. 98, No. 1 Weinstock

rate constants and Ea values in Table 1 are small and


vary with ionic strength.)
Irrespective of the explanation for the low charge
product calculated from Figure 1, the Coulombic work
term at infinite dilution should reflect the ideal (no
ion pairing) charge product of z1 z 2 = 30. Hence, the
slope of the plot of ln k vs exp(–χr 12) should have
steepened as the ionic strength approached 0 M.
However, there is no way of knowing at what ionic
strength this might have occurred. This uncertainty
makes extrapolation of k to infinite dilution prob-
lematic. The value of k. = 4.5 × 10–3 M–1 s–1
reported by Rasmussen and Brubaker was obtained
by extrapolating to µ = 0 M, using the slope of Ln k
vs exp(–χr12) shown in Figure 1 (slope = z1 z2 = 7.8).
Values of k obtained from the data of Rasmussen
and Brubaker have been used to determine the self-
exchange reorganization energy, λ11, using eq 47
Table 1. Dependence of Rate Constant on
(from section II.B.3):
Temperature and Ionic Strength

where W(r) is given by

The k0 value reported by Rasmussen and Brubaker


was obtained by extrapolation of the line plotted in
Figure 1 to infinite dilution (exp(–χr12) = 1). Thus,
used in the experiments. 79 This error raises a serious the k0 value they reported could only have been
question regarding the reliability of their data for use correct if the slope of the plot shown in Figure 1 (z 1 z2
in the calculation of fundamental theoretical param- = 7.8) had been constant to µ = 0 M. Thus, in
eters. choosing Rasmussen and Brubaker’s value of k0 as a
In addition, there is some suggestion from the work basis for calculating λ 11, one is forced to use a charge
of Kozik and Baker that ion pairing with Li+ cations product of 7.8 for calculation of the Coulombic term,
in the [COIIIW12O40]5–/[CoIIW12O40]6– system would not W(r), in eq 49. Doing so (r = 11.2 Å, T = 273 K), one
occur to a significant extent at moderate ionic- obtains, from eq 47, λ = 61.2 kcal mol–1 (Z = 1011
strength values. The evidence is from a study of the M–1 s–1) or λ = 56.2 kcal mol–1 (Z = 1012 M–1 s–1).
α-[P2W18O62]6–/ α-[P2W18O62]7– self-exchange reac- A formally analogous method was used by Amjad,
tion.23 Experimental rate-constant values for this Brodovitch, and McAuley in 1977.76 Returning again
reaction were 3 orders of magnitude smaller than to Figure 1, Amjad et al. extrapolated to an ionic
calculated ones, even at fairly high ionic strengths strength of µ = 1.0 M, which was pertinent to their
(µ = 1.49 M, LiCl). Localization of the blue electron study of the oxidation of organic acids and pheno-
notwithstanding, if ion pairing were occurring, one lic compounds by [COIIIW12O40]5–. At µ = 1.0 M,
would expect observed rate constants to have ex- exp(–χr 12) = 0.037; the contribution of Coulombic
ceeded, rather than to have fallen short of, calculated work to ∆G ‡ was less than 4% of that at µ = 0 (eq 1).
values. Hence, the contribution of W(r) exp(–χr 12) to ∆G ‡ was
Finally, Rasmussen and Brubaker measured the ignored. The free energy of activation, ∆G ‡, calcu-
temperature dependence of k as a function of ionic lated from the rate constant obtained by extrapola-
strength (Table 1). tion to µ = 1.0 M (k( µ = 1.0M) = 1.7 M–1 s–1), was set
These data were used to extrapolate to infinite equal to λ/4 (eq 14). A value of λ = 4(15.36 kcal
dilution, which gave Ea (µ =0M) = ~18 kcal mol–1. Using mol–1 = 61.4 kcal mol–1 was obtained.
a preexponential factor of Z = 1011 M–1 s–1, this value Fair agreement between theory and experiment
of Ea would give (using eq 4) k0 = 1011 exp(– Ea /R T) was found when the extrapolated self-exchange rate-
= 3.9 × 10–4 M–1 s–1 at 0 °C. constant value (k(µ=1.0M) = 1.7 M–1 s–1 was used to
Unfortunately, a conceptual error was made in predict the rate constant, k12, for cross-reaction
Rasmussen and Brubaker’s extrapolation µ = 0 M. between CoIIIW125– and ascorbate anion.76,80 In ad-
To determine Ea (µ =0M) from data of this kind, it is dition, [COIIIW12O40]5– was reacted with [Ru(CN)6]3–
necessary to determine k0, by extrapolation to µ = 0 and the Marcus cross-relation (eq 17) used to calcu-
M, at each temperature studied.19 (Note that the late the self-exchange rate constant, kRu-Ru, for the
Arrhenius preexponential factors indicated by the Ru(CN)63–/4– couple. Notably, the kRu-Ru value cal-
Electron-Transfer Reactions of POMs Chemical Reviews, 1998, Vol. 98, No. 1 127

Table 2. Kinetic and Thermodynamic Parameters of Redox Reactions Involving POM Anions in Water at 298 K;
0.5 M HClO 4, µ = 1.0 M (NaClO4) a

culated using the extrapolated rate-constant value


(1.7 M–1 s–1) was consistent with kRu-Ru values
determined by other cross-reactions of the Ru com-
plex.81 These results might suggest that the reorga-
nization energy of the CoIIIW125–/CoIIW126– electron
self-exchange reaction is near 60 kcal mol–1.
In 1987 Kozhevnikov78 revisited the data reported
by Rasmussen and Brubaker and pointed out (foot-
note to reaction 11 from Table 1, p 481 in the English
translation of the cited article) that the experimental
k0 value originally reported was likely in error (too
large) by several orders of magnitude. Rather than
use the slope of the data shown in Figure 1, he
extrapolated to µ = 0 using a charge product of z 1 z 2
= 30 (no other details were provided). This gave an
experimental value of k0 = 5.03 × 10–6 M–1 s–1, from
which λ11 values of 46.1 (Z = 1012 M–1 s–1, T = 273
K) or 41.1 (Z = 1011 M–1 s–1) are obtained using eqs
47 and 49.
In addition, Kozhevnikov reported rate constants,
activation parameters, and thermodynamic data for
cross-reactions between POM anions and between
POM anions and organic substrates or other inor-
ganic complexes (Table 2). From eq 6
the same set of reactions, λ = 78.9 kcal mol–1, which
appears to be in error by a constant factor (calcula-
tion of λ from eq 7, as it appears on page 482 of the
English edition of the cited article,78 gives a value of
Reorganization energies, calculated by solving eq 50 81 kcal mol–1).
for each reaction, are included in the rightmost Attributing the scatter in Figure 4 to uncertainty
column of Table 2. in k0 values, Kozhevnikov argued that the POM
Values of RT ln k0 + W(r), as a function of ∆G°', anions used in reactions 1-17 possessed a common
are plotted as shaded diamonds in Figure 4. A inner-sphere reorganization energy. On this basis,
nonlinear regression of the data (solid line, Z = 1011 it was concluded that the barrier to reduction of the
M–1 s–1) was performed using the Solver program in Co(III) ion in [CoIIIW12O40]5– entailed no additional
Microsoft Excel by minimizing the sum-of-error- energy over that involved in reduction of the ad-
squared with respect to the reorganization energy, dendum atoms–V(V), Mo(VI), or W(VI)–of Keggin
1. Using a preexponential factor of Z = 1011 M–1 s–1, anions possessing PO or Si(IV) heteroatoms. Fur-
the best fit was found for λ = 38.6 kcal mol–1 (λ = thermore, the absence of an additional energy barrier
42.6 kcal mol–1 for Z = 5 × 1011 M–1 s–1 and 44.4 was taken as evidence that contrary to the hypothesis
kcal mol–1 for Z = 1012 M–1 s–1). These values differ of Varga, Papaconstantinou, and Pope (summarized
substantially from that reported by Kozhevnikov78 for above), the central atom in [CoIIIW12O40]5– was not
128 Chemical Reviews, 1998, Vol. 98, No. 1 Weinstock

“electronically isolated” from the tungsten frame- tion energies of the POM anions. In addition, ob-
work. Some of the data used in reaching this serving no additional energy barriers to electron
conclusion warrant further scrutiny. transfer, Kozhevnikov was led to suggest that CoO4
The reactions between [SiW12O40]5– and [PW12O40]4– moieties present at the center of α-[CoW12O40]5–/6–
and Fe(CN)6]3– (reactions 5 and 6 in Table 2) were anions are not electronically isolated from the W12O36
carried out at pH 1 in 0.1 M HClO4 (vs 0.5 M HClO4 shell. However, convincing data obtained from ESR
used for other reactions). Kozhevnikov calculated the and NMR spectroscopies, measurements of magnetic
standard free energy of the reactions using the properties, and X-ray crystallography30,64,88 suggest
commonly published value of the standard reduction otherwise.
potential for the [Fe(CN)6]3–/(Fe(CN)6]4– couple, 0.36 Eberson used the k0 value that Rasmussen and
V vs the normal hydrogen electrode (NHE). How- Brubaker calculated based upon a charge product of
ever, this reduction potential value applies only to z1 z2 = 7.8 (Figure 1). At the same time, however,
neutral or basic [Fe(CN)6]3– solutions; at pH values Eberson chose the more intuitively compelling, yet
between 3 and 4, [HFe(CN)6]3– rather than [Fe(CN)6]4– internally inconsistent, charge product of z1 z2 = 30
is the dominant species in solution, and at pH 2, for calculation of the work term at infinite dilution
[H2Fe(CN)6]2– predominates.82–84 Hence, at pH 1 in (i.e., in eq 1 with exp(–χr 12) = 1).59,77 Taking Z =
0.1 M HClO4, the reduction potential of the [Fe- 1011, Eberson calculated λ = 23 kcal mol–1 at 0 °C.
(CN6]3–/[Fe(CN)6]4– couple is actually ~200 mV This reorganization energy value, which is very close
(~4.6 kcal mol–1) more positive than that at neutral to that determined for the α-[PW12O40]3–/α-[PW12O40]4–
pH. In entries 9 and 10 (Table 2), [PV25–/6– W10O40]5–/6– system (λtheoretical = λin + λout = 16.4 + 6.1= 22.5 kcal
anions were reacted with [PV2Mo10O40] anions. mol–1 at 25 °C), appears to be too small.
Although attempts were made to evaluate the extent Furthermore, in 1992 Eberson77 used Kozhevnik-
of vanadium dissociation from the more labile vana- ov’s data to recalculate a common reorganization
domolybdophosphate anion and the impact this dis- energy of ~28 kcal mol–1 for electron-transfer re-
sociation might have had on the rate of electron actions involving the anions [PW12O40]3–/4–
transfer,85 the kinetic data reported for these reac- [SiW12O40]4–/5–, and [CoW12O40]5–/6– (entries 1, 2, and
tions must be viewed with caution. Additionally 11, Table 2). The calculation was based on the
complicating these reactions is the tendency of thermodynamic and kinetic parameters compiled by
[PV2Mo10O40]5– anions to exist in solution not only Kozhevnikov.78 However, recent recalculation by
as mixtures of positional isomers, but as equilibrium Professor Eberson89 gives values of λ11 = 49, 48, and
mixtures86,87 of anions: [PVMo11O40]4–, [PV2Mo10O40]5–, 43 kcal mol–1, respectively, for these reactions (1, 2,
and [PV3Mo9O40]6–. In reactions 16 and 17, reduced and 11). Moreover, it is likely that the reorganization
POM anions were reacted with O2. Complications energy of the [CoW12O40]5–/6– self-exchange reaction
inherent in evaluating the reorganization energy is considerably larger than that for [PW12O40]3–/4– or
associated with the reduction of O2 to O2•– are [SiW12O40]4–/5– (see section III.C, above). If so, the
discussed in section VIII. reorganization energy for the [CoW12O40]5–/6– couple
Calculated reorganization energies (Table 2, Z = could be significantly greater than 43 kcal mol–1 (λ 1 2
5 × 1011) for the reactions of [CoIIIW12O40]5– with well- = ½(λ 11 + λ22), eq 43).
defined POM anions (entries 1 and 2 in Table 2 and In summary, it appears that the reorganization
labeled 1 and 2 in Figure 4) and with [CoIIW12O40]6– energy for the CoW125–/6– self-exchange reaction may
(self-exchange; entry 11 in Table 2 and labeled 11 in be substantially higher than that of the α-(PW12)3–/4–
Figure 4) were 54, 52, and 47 kcal mol–1, respectively. couple or related heteropolyanion/heteropoly blue
TheIII/II
value of 47
5–/6–
kcal mol–1 for self-exchange of the pairs. One way to see this directly is to compare the
[Co W12O40] couple followed from the k0 value rate constant for the CoW125–/6– self-exchange reac-
Kozhevnikov obtained by extrapolation from the data tion with that of the Wells-Dawson anions (α-
of Rasmussen and Brubaker, using z1 z2 = 30, as [P2W18O62]6– and α-[P2W18O62]7–). The rate constant
noted above. Significantly, the reorganization energy observed for self-exchange between the Wells-Daw-
values for the cross-reactions between α-[SiW12O40]5–, son anions at µ = 0.59 M was 1.1 × 103 M–1 s–1
or α-[PW12O40]4–, and α-[CoIIIW12O40]5– (reactions 1 compared to 0.64 M–1 s–1 for the CoW125–/6–couple
and 2) are quite high (54 and 52 kcal mol–1). Using at a similar ionic strength of 0.61 M. Even though
the reorganization energy calculated by Kozik and the Wells-Dawson anion has a higher charge and
Baker for the α-[PW12O40]3–/4– self-exchange reaction thus a larger value of W(r) at a given ionic strength,
(23 kcal mol–1), these data give reorganization ener- electron self-exchange is much faster than for CoW12.
gies for the [CoIII/IIW12O40]5–/6– self-exchange reaction The difference in rates is likely attributable to a
of 85 or 81 kcal mol–1, which seem a bit high and higher value for the inner-sphere reorganization of
may reflect errors arising from extrapolations to k0 . the Co ion.
In summary, Kozhevnikov’s suggestion that the In their original paper, Rasmussen and Brubaker5
POM anions listed in Table 2 might possess similar calculated a λtheoretical value for the CoW125–/6– couple
reorganization energies is unwarranted. This sug- of λin + λout = 11.1 + 18.4 = 29.5 kcal mol–1 at 0 °C.
gestion was reached by attributing the scatter in The outer-sphere reorganization energy is very simi-
Figure 4 to experimental uncertainties. However, in lar to that calculated by Kozik and Baker,23 the
light of the more recent work of Kozik and Baker, it discrepancy likely due to differences in temperature
is now clear that the scatter in Figure 4 can be and in accepted ionic radii for the Keggin anion. The
attributed to real differences between the reorganiza- larger value of λin calculated for the [CoIIIW12O40]5–/
Electron-Transfer Reactions of POMs Chemical Reviews, 1998, Vol. 98, No. 1 129

[CoIIW12O40]6– system (Kozik and Baker estimated λi n Using 0.41 V vs NHS for the reduction potential
= 6.1 kcal mol–1 for the α-(PW12)3–/4– couple) arose of the Fe(CN)6]3–/Fe(CN)6]4– couple at pH 4.5 and
from consideration of the difference in Co-O bond –0.02 V for the [CuIIW12O40]6–[CuIW12O40]7– couple
lengths between tetrahedral CoIIIO4 and CoIIO4. at the pH and ionic strength of the cross-reaction, a
However, their calculation did not account for distor- rate constant of k11 = 2.61 M–1 s–1 was calculated
tion seen in the X-ray crystal structure90 of the for self-exchange between the copper complexes. The
[CoIIIW12O40]5– anion, which was subsequently con- purpose of this study was to investigate the nature
firmed by accurate magnetic measurements63,64 and of electron transfer in blue copper proteins, such as
by variable-temperature 183W NMR spectroscopy91 to plastocyanine and azurin, in which both Cu(I) and
be Jahn–Teller distorted. The CoIIO4 moiety in Cu(II) oxidation states share a common distorted
[CoIIW12O40]6–, on the other hand,64,91 was found to tetrahedral geometry. The self-exchange rate con-
describe a near-perfect tetrahedron. Hence, one stant, which had been calculated for the plastocyanin
explanation for the slow rate of self-exchange in the system, was ca. 105 M–1 s–1; it was thought that a
[CoIIIW12O40]5–/[CoIIW12O40]6– system is that a large similar rate might be observed in the [CuW12O40]6–/7–
inner-sphere reorganization energy is associated with system, the central Cu ions similarly residing in
reduction of the CoO4 moiety. Kozik, Hammer, and pseudotetrahedral environments. The surprisingly
Baker point out that the much slower rate is “not small self-exchange rate constant found for the
surprising because...of the considerable energy bar- [CuW12O40]6–/7– system was attributed to the need for
rier involved in interconversion of the small Jahn– the transferring electron to pass through the W12O36
Teller distorted CoIIIO4 tetrahedron and larger regu- shell, a proposal originally set forth by Rasmussen
lar CoIIO4 tetrahedron”.25 and Brubaker to explain the slow rate of electron
In light of the well-established electronic isolation exchange in the [CoW12O40]5–/6– system.5
of the central Co(II) ion in [COIIW12O40]6–, poor However, unlike in the electronic spectra of
electronic overlap between the d orbitals of the Co- [CoW12O40]6–, some intervalence charge transfer (i.e.,
(II) and Co(III) ions may also result in low values for Cu(I) f W(VI) = Cu(II) f W(V)) has been observed
vel (i.e., transmission coefficients less than unity). in [CuIW12O40].7–94 Thus, as observed in the self-
Poor electronic overlap has been suggested as an exchange of [Ru(bpy)3]2+/3+,53 it might be argued that
explanation for slow rates of electron transfer ob- electron delocalization of this type would allow for
served for a number of CoII/CoIII reactions.53,92 In greater electronic overlap between the Cu(I) and Cu-
octahedral Co(III) complexes, spin-multiplicity re- (II) d orbitals in the [CuW12O40]6–/7– system, thus
strictions have also been suggested as an explanation minimizing the insulating effect of the W12O36 shell.
for slow rates of exchange. However, analysis has At the same time, by analogy with the considerable
shown that coupling of the vibrational modes of the distortion seen for the CoIIIO4 moiety in [CoW12O40]5–,
ground electronic states is sufficient to lift the spin- more energy might be required for the Cu ions to
state restrictions, such that when vibronic coupling obtain their favored geometries (square planar for
is considered, the reduction of Co(III) is in fact Cu(II) and tetrahedral for Cu(I)) in the [CuW12O40]6–/7–
allowed.92 Although no spin-state restrictions would anions than in the enzymatic system. Hence, the low
apply to the reduction of tetrahedrally coordinated value for self-exchange between [CuW12O40]6–/7– an-
Co(III), the situation with respect to Jahn–Teller ions might also be attributed to large inner-sphere
distorted CoIIIO4 in [CoIIIW12O40]5– may be different. reorganization energies and not to “electronic isola-
Finally, much of the above discussion has assumed tion”, as the literature might lead one to conclude.
that, despite flaws in their published data, the low
rates of electron self-exchange between [CoIIIW12O40]5– IV. Oxidation of Organic EIectron Donors by POM
and [CoIIW12O40]6– measured by Rasmussen and Anions
Brubaker were reliable. If so, it would be untenable
to argue that the reorganization energy of the A. Alkylaromatic Compounds
CoIII/IIW125–/6– couple is not significantly larger than
that of the PW124–/3– couple. Unfortunately, in light In 1970 Chester recognized the potential utility of
of the flawed nature of some of their data, the ko b s [CoIIIW12O40]5– (CoIIIW125–) as a well-defined probe for
value published by Rasmussen and Brubaker may determining the nature of outer-sphere oxidations of
not be reliable. alkylaromatic hydrocarbons.7 Neat toluene and o-,
m-, and p-xylene were reacted with K5[CoIIIW12O40]•-
D. One-Electron Oxidation of [Cu I W 12 O 40 ] 7– to 1H2O under heterogeneous conditions at 96 °C.
[Cu I I W 1 2 O 4 0 ] 6 – Analogous reactions were carried out in refluxing.
Slow rates were also observed for electron transfer aqueous acetic acid (15 mL each of water and the
at tetrahedral copper in the room-temperature reac- organic substrate and 75 mL of acetic acid). The
tion between [CulW12O40];- and [Fe(CN)6]3– at pH 4.5 products of the heterogeneous oxidations were domi-
in acetate buffer (ionic strength adjusted to 0.1 M nated by isomeric diphenylmethane derivatives, which
using NaClO4).93 The reaction is first order in implicated the intermediacy of benzylic cations.
[CuIW12O40]7– and Fe(CN)6]3–, and a second-order Oxidation of toluene in the presence of excess ben-
rate constant of k12 = 4.1 × 105 M–1 s–1 was zene gave diphenylmethane, which is consistent with
measured. The Marcus cross-relation (eqs 17 and 18, electrophilic aromatic substitution of benzene by
Z = 1011 M–1 s–1) was used to calculate the rate benzyl cation intermediates. In aqueous acetic acid,
constant for the [CuIIW12O40]6–/[CuIW12O40]7– self- benzyl cation intermediates gave rise to benzyl
exchange reaction. acetates.
130 Chemical Reviews, 1998, Vol. 98, No. 1 Weinstock
II 5– –
Because Co W12 was a known outer-sphere electrons, probably irreversibly; the 2-e -reduced
oxidant, oxidation of the alkylaromatic compounds anions [NiIIMo9O32]8– and [MnIIMo9O32]8– have not
to benzyl radical cations, followed by expulsion of a been reported. These complications mitigated against
proton, was seen as a more likely route to intermedi- use of the nonamolybdometalate complexes for fur-
ate benzylic radicals than was direct abstraction of ther mechanistic study of the alkylbenzene oxidation
a hydrogen atom, H•. The absence of dibenzyl reactions.
products, which form rapidly via coupling of benzylic Following the observation of a kinetic-isotope effect
radicals, suggested that the benzyl radical intermedi- in the oxidation of PMT by CoIIIW125–, Eberson
ates were preferentially oxidized to cationic species carried out a detailed investigation that clarified the
by an additional equivalent of CoIIIW125–. No kinetic role of proton and electron transfer in this reaction
studies were performed. However, it was proposed and laid the groundwork for the subsequent use of
that under heterogeneous conditions, oxidation oc- CoIIIW125– as a probe for the study of outer-sphere
curred via overlap of the aromatic π system with the electron-transfer processes.59 CoIIIW125– was reacted
tungstate framework “conduction bands”. Electron with PMT in 55:45 (w/w) acetic acid to water. The
transfer to the W12O36 framework was followed by water was included to keep the dielectric constant of
rapid reduction of the Co(III) ion. An analogous the system as high as possible in order to avoid ion
mechanism was proposed for the homogeneous oxi- pairing. The reaction was followed at 50 °C by
dation of alkylaromatics by CoIIIW125– in aqueous monitoring the absorbance of the CoIIIW125– anion
acetic acid.7 Despite the potential intermediacy of a (λmax = 390 nm) in the presence of large (100-fold)
charge-transfer complex between substrate and the excesses of PMT. A nonlinear regression method was
[CoW12O40]5– complex, the Co(III) ion, buried in the used to determine the fit of proposed rate expressions
center of the [W12O40]8– framework, was entirely to absorbance vs time data. In addition to kinetic
inaccessible by the substrate. It was thus concluded isotope effects, the effects of added base (AcO–); of
that direct coordination of alkyl aromatics to metal alkali-metal, alkaline-earth, and tetraalkylammo-
cation oxidants, in this case Co(III), was not a nium cations; and of the dielectric constant of the
necessary prerequisite to homogeneous oxidation. medium were also determined under conditions of
In 1980 Eberson published the first in a series of constant ionic strength. Prior to theoretical analysis
papers that characterized the homogeneous oxida- of the data (Marcus), the reduction potential of
tions of alkylaromatic compounds by POM an- CoIIIW125– was determined in the reaction medium
ions.8,59,77,95 A preliminary study of the oxidation of (including added electrolytes) used for the acquisition
p-methoxy toluene (PMT) and of its deuterated of kinetic data.
derivative, 4-methoxy(α,α,α-2H3)toluene, revealed a After failure to fit the experimental data to expres-
kinetic-isotope effect of kH /kD = ~6–clear evidence sions of absorbance vs time derived by integration of
for proton transfer in the rate-determining step. The simple rate laws, an excellent fit was found for an
oxidations of toluene, xylenes, and mesitylene were expression derived by integration of a rate law that
also studied, although kinetic-isotope data were not described the elementary steps shown in eqs 51 and
obtained. Because CoIIIW125– was considered an 52 (B is either H2O or AcO–).
outer-sphere oxidant, the observation of rate-limiting
proton transfer led Eberson to propose that electron-
transfer oxidations of C-H bonds might involve a
“wide spectrum of electron transfer/proton-transfer
mechanisms, ranging from ‘pure’ outer-sphere...to an
inner-sphere mechanism in which an H atom is
transferred from the C–H bond to a ligand atom of A steady-state approximation (d[PMT• +]/dt = 0)
a metal complex”.8 Shortly thereafter, the notion of was used to derive a rate expression for the disap-
simultaneous electron and proton transfer was de- pearance of CoIIIW125– (eq 53).
scribed theoretically by Perrin.96 A similar range of
mechanisms were observed independently by Meyer
in oxidations of C-H bonds by polypyridyl complexes
of ruthenium and osmimn.97
Jönsson98 studied the oxidation of alkylaromatic
hydrocarbons by [NiIVMo9O32]6– and [MnIVMo9O32]6–99
in aqueous acetic acid. The ratio of acetoxylated (4- Substitution of [CoIIW126–] = Ao – A, where A o, is the
methoxybenzyl acetate) to hydroxylated (4-methoxy- initial absorbance of CoIIIW125– and A its absorbance
benzyl alcohol) products increased with increase in at time = t, into eq 53, followed by integration, gave
the acetic acid-to-water ratio of the solvent. By an expression for absorbance vs time, which by
analogy with the reactions of these substrates with nonlinear regression, provided an excellent fit to
CoIIIW125–, the reactions were assumed to proceed via experimental data. The proposed rate expression
solvolysis of benzylic cation intermediates. Notably, was further verified by systematic variation of key
kinetic isotope effects of kH /kD = 4.2 and 2.7 were variables, [PMT], [B], and [CoIIW126–], in eq 53.
observed in the oxidations of toluene and 4-methoxy- The data were consistent with a mechanism in
toluene by [NiIVMo9O32]6–. Unlike Co(III) in CoIIIW125–, which a rapid and reversible electron-transfer step
the Ni(IV) and Mn(IV) heteroatoms in [NiIVMo9O32]6– (eq 51) was followed by rate-limiting reaction of
and [MnIVMo9O32]6– are readily reduced by two PMT • + with base to give PMT• . The benzylic radical,
Electron-Transfer Reactions of POMs Chemical Reviews, 1998, Vol. 98, No. 1 131

PMT•, was rapidly oxidized by a second equivalent reduction potential (i.e., with less easily oxidized
of CoIIIW125– to give a cation that reacted with acetic neutral substrates). Hence, one way to observe rate-
acid (eqs 54 and 55). limiting electron transfer, it was argued, might be
to use alkylaromatic substrates less easily oxidized
than PMT. To oxidize these compounds, POM anions
with reduction potentials more positive than that of
CoIIIW125– would be required. Investigation in this
direction led not to more direct measurement of the
rate of outer-sphere electron transfer (k 1), but to the
Additional evidence for rate-limiting proton trans- observation of synchronous proton and electron trans-
fer was provided by the observation of a kinetic fer.
isotope effect of -6 when the reaction was carried p-Xylene (irreversible anodic potential, Eanodic =
out using 4-methoxy(a,a,a-2H3)toluene. At low con- 2.18 V vs NHE in acetonitrile, compared with 1.82
centrations of acetate ion, such that B = H2O (the V for PMT) was reacted with (NH4)6H2[CeIVMo12O42]
concentration of water being ~25 M), k2 was found in aqueous acetic acid. 95 In 3:l water to acetic acid,
to be 80 M–1 s–1, while when [AcO–] was increased five anodic waves were observed in the cyclic volta-
to 60.5 M, such that B = AcO–, k2 increased to 3.6 ×
–1 –1 mmogram of (NH4)6H2[CeIVMo12O42] (graphite elec-
10 M s . trode, no added electrolyte). The most positive of
Values for k1 and k–1 were calculated (eqs 1 and 4) these, an irreversible anodic wave observed at -1.42
using values for the reorganization energy of the V (NHE), was assigned to the oxidation of Ce(III) to
CoIIIW125–/CoIIW126– self-exchange obtained from the Ce(IV). The next most positive wave was observed
data of Rasmussen and Brubaker (λ11 = 16-23 kcal at ~1.2 V. Eberson assigned the couples observed
mol–1 at 50 °C), and λ22 = 9.0 kcal mol–1 for the PMT/ at 1.2 V and at more negative potentials to the
PMT•+ self-exchange reaction. The reorganization reduction of Mo(VI) ions in [CeIVMo12O42]8–. These
energy, λ12, for reaction of CoIIIW125– with PMT was redox couples were reported to be quasi-reversible or
then calculated: λ12 = (½)(λ11 + λ22) = 16 kcal mol–1. nearly so (Eanodic – Ecathodic = 100, 55, 120, and 65 mV
As discussed above (section III.C), the reorganization at sweep rate of 50 mV s–1). No distinct cathodic
energy for the CoIIIW125–/CoIIW126– couple is probably peak, which would have corresponded to the reduc-
somewhat higher, and the λ value used for the PMT/ tion of Ce(IV) to Ce(III), was seen. The absence of
PMT •+ self-exchange reaction may also have been this peak was attributed to the “reduction” of water.
underestimated.77 However, analysis by the author It is unclear how water could be reduced at such
showed that error in estimation of reorganization positive electrode potentials (>+1.3 V).
energies did not greatly affect the calculated ratio,
k– 1 /k1, which for a reorganization energy of 16 kcal In polarographic analysis of the [CeIVMo12O42]8–
mol–1, was found to be ~1010. Greater uncertainty anion in 1.0 M sulfuric acid, Pope100 observed revers-
was attributed to the effects of ion pairing on the ible reduction of Ce(IV) to Ce(III) (+0.68 V). The
forward reaction rate constant, k1 . discrepancy between this value and that reported by
A specific-cation effect was observed by replacing Eberson appears to exceed the range of values
Na+ by a series of alkali-metal, alkaline-earth, defined by study of medium (solvent and electrolyte)
and tetrabutylammonium cations at constant ionic effects on the reduction potential of the CoIII/IIW125–/6–
strength. The observed rate constant decreased in couple.52 No further reduction was seen by Pope
the order Sr2+ > Cs+ > Ca2+ > Rb+ > K+ > Li+ > until the electrode potential reached –0.1 V, at which
Na+ > Me4N+ > Pr4N+ > Hex4N+ > Dec4N+. Ca- a complex multielectron process was observed. The
talysis by H+ was also noted. These results sug- absence of an anodic peak after multielectron reduc-
gested that significant ion pairing was occurring. The tion of [CeIVMo12O42]8– was consistent with irrevers-
charge products z1 z2, used to calculate Coulombic ible reduction of the anion. (Irreversible reduction of
terms and to correct the standard free energy of addendum atoms is typical of “type II” polyanions, such
reaction (eq 1, with z1 = –5 for CoIIIW125– and z2 = as [CeIVMo12042]8–, in which MO6 octahedra-where
–6 for CoIIW126–), were probably representative of M is an addendum atom, usually Mo(VI) or W(VI)–
limiting rather than prevalent values. Although the possess two cis terminal oxygens.)101 However, apart
data did not allow for strict quantification of the from Eberson’s seemingly high reduction potential
effects of ion pairing on observed rate constants, value for the l-e- reduction of [CeIVMo12O42]8–, which
approximate equilibrium constants for ion pairing of when used to calculate the standard free energy of
CoIIIW125– with Na+ (K 1 = ~101 M–1) and with Bu4N+ reaction with substrate, would have a significant
(K 2 = 103 M–1) were estimated (the above trend in impact on theoretical analysis, the use of this poly-
specific-cation effects may reflect greater extents of anion as a reversible-potentially outer-sphere-single
ion pairing with CoIIW126–, vs with CoIIIW125–, which electron oxidant appears sound.
would decrease the ratio k1 /k–1 in eq 51). The products of oxidation of p-xylene by
The results of this study suggested that if the rate [CeIVMo12O42]8– (CeIVMo128–) were the ones (4-meth-
of proton abstraction from the cation-radical inter- ylbenzyl acetate and 4-methylbenzyl alcohol) ex-
mediate (k2 in eq 52) could be increased, outer-sphere pected to result from the formation of benzyl cation
electron transfer might become rate limiting and the intermediates under the prevailing reaction condi-
forward rate constant for electron transfer (k 1 in eq tions. Progress of reaction was followed by observa-
51) observed directly. 95 It was known that rates of tion of the absorbance of the [CeIIIMo12O42l9– anion
C-H bond cleavage in cation radicals increased with at 410 nm. At 50 °C under pseudo-first-order condi-
132 Chemical Reviews, 1998, Vol. 98, No. 1 Weinstock

tions ([ArCH3]0/[CeIVMo] > 37.5), absorbance vs time Table 3. Summary of Self-Exchange Reorganization
data were fitted to an expression derived by integra- Energies Calculated from Crow-Reactions with
Inorganic and Organic Oxidants
tion of the second-order rate law shown in eq 56.

A value for k1 of 9.4 × 10–2 M–1 s–1 was determined,


and a kinetic isotope effect of 1.5 was observed. In
contrast to the reaction between CoIIIW125– and
PMT, the present reaction was zero order in
[(CeIIIMo12O429–)] and in [AcO–]. Despite the high
value assigned to the reduction potential of Ce(IV)
in [CeIVMo12O42]8–, theoretical analysis of the reac-
tion, assuming a rate-limiting outer-sphere electron-
transfer step, gave rate constants 5–6 orders of mag-
nitude smaller than observed values. Uncertainties
with regard to the reorganization energy of the [(CeIV-
Mo12O42]8–/[CeIIIMo12O42]9– system, the extent of ion
pairing present, the pK a of the [CeIIIMo12O42]9– anion, enough to keep the probability of electron transfer
and related issues were acknowledged and system- occurring, once Franck-Condon requirements have
atically discussed. After careful consideration, it was been satisfied, equal to unity). The inner-sphere
concluded that agreement with Marcus theory could step, case 2, can be assumed to embody an electronic
not be obtained without unreasonable assumptions. interaction of near 5 kcal mol–1 or greater, while case
Finally, it was suggested that the reaction might not 3, a polar pathway, is likely associated with an
proceed by simple outer-sphere electron transfer and electronic interaction of 15–20 kcal mol–1. (Charge-
that an alternative mechanism, one involving syn- transfer complexes, often associated with electronic
chronous electron-proton transfer, should be seri- interactions of 1–2 kcal mol–1, may be viewed as
ously considered. forming en route to an inner-sphere electron-transfer
All the same, the authors argued that the relatively step, case 2.)
Weak kinetic isotope effect (kH /kD = 1.5) did not In previous work, Eberson, in collaboration with
unequivocally exclude a rate-limiting outer-sphere Shaik,103 had performed a theoretical and experi-
electron-transfer step. The kH /k D value of I.5 might mental analysis of the organic electron-transfer reac-
have been due to secondary kinetic isotope effects, tions of organic radical anions. The authors con-
including the effect of deuterium substitution on cluded that in reactions of radical anions, RH• –, with
inner-sphere reorganization energies.77,95,102 neutral organic molecules, RH, sizable electronic
Nonetheless, the possibility of concerted electron- interactions (~3–8 kcal mol–1) were involved in the
proton transfer was strongly supported by the unex- relevant transition states. Similar interactions were
pectedly rapid rate observed for the oxidation of predicted for reactions of radical cations, RH• + .
p-nitrotoluene (E° > 3 V) by CoIIIW125–.8,59 Given the Cross-reactions between POM anions and neutral
high endergonicity associated with oxidation of p- organic compounds, RH, were used to estimate
nitrotoluene to its radical cation, the rapid oxidation reorganization energies, λ, associated with electron
of p-nitrotoluene by CoIIIW125– argued against an self-exchange between RH and l-e--oxidized deriva-
outer-sphere electron-transfer mechanism. Con- tives, RH•+. The estimated reorganization energies
certed electron-proton transfer was suggested. were then used to calculate (Marcus) the rate con-
In 1992 Eberson77 used POM anions and other stants expected for outer-sphere electron transfer
outer-sphere oxidants to clarify fundamental issues between RH and RH• +. Differences between theo-
concerning the mechanisms of reactions involving the retically and experimentally determined rate con-
formation of radical cations from organic substrates. stants indicated that electronic interactions in the
Three pathways for electron transfer between organic transition state were greater than those allowed for
donor, D, and acceptor, A, each involving greater by an outer-sphere mechanism.
degrees of electronic coupling in the transition state, In addition to POM anions, data obtained using a
were considered (Scheme 3). number of other inorganic outer-sphere oxidants were
reviewed (entries 1–6, Table 3). The reorganization
Scheme 3 energies listed for entries 7–10 were obtained by
Marcus treatment of kinetic data obtained from the
indicated self-exchange reactions.104
The reorganization energies calculated from rate
data for the self-exchange reactions (entries 7–10,
Table 3) were consistently lower than those calcu-
lated from rate data obtained from the cross-reactions
with inorganic outer-sphere oxidants. Although no
rate data for self-exchange of alklyaromatic com-
Case 1, outer-sphere electron transfer, is defined pounds, such as those referred to in entries 1–5, were
by the criterion that electronic coupling in the transi- available, the difference between the λ values calcu-
tion state is less than ~1 kcal mol–1 (but still large lated for the reaction of naphthalene with OsCl6– and
Electron-Transfer Reactions of POMs Chemical Reviews, 1998, Vol. 98, NO. 1 133

the λ values calculated for the self-exchange reactions Table 4. Kinetic Data for the Oxidation of ArCH 2R by
between the polycyclic aromatic compounds listed in [CoIIIW 12O40]5–
entries 7 and 8, suggested that the λ values obtained
by Marcus treatment of the self-exchange reactions
were low by ~20 kcal mol–1. Because λ values
contribute λ/4 kcal mol–1 to the activation energies
of outer-sphere electron-transfer reactions (eq 11, the
20 kcal mol–1 error would correspond to an electronic
interaction of ca. 5 kcal mol–1 in the transition states
of the ArH/ArH• + self-exchange reactions (case 2,
Scheme 3). This result, consistent with those ob-
tained in analysis of self-exchange between ArH and
radical anions, ArH• –,103 led to the conclusion that
outer-sphere mechanisms may be less kinetically It was anticipated that careful kinetic analysis would
facile than inner-sphere pathways in electron-trans- provide information regarding the role of the a-sub
fer reactions between neutral organic compounds and stituents, R, in the two kinetically pertinent steps of
cationic or anionic organic radicals. the oxidation process (eqs 57 and 58).
The reorganization energy determined for the ArH/ Reactions of CoIIIW125– with ArCH2R were carried
ArH• + self-exchange reaction reported in entry 1 out at 50 °C in the presence of 0.47 M KOAc. Under
(Table 3, λ = 55 kcal mol–1 was calculated using a these conditions, side-chain-substituted α-hydroxy
reorganization energy (λ11) for the CoIII/IIW125–/6– and α-acetoxy products (ArCXHR, X = OH and OAc)
couple of 23 kcal mol–1.77 Cause for concern regard- were formed when R = H or Me. With R = OH, OMe,
ing the accuracy of this λ11 value, which was calcu- or OAc, however, the first-formed α-hydroxy or ac-
lated from the data of Rasmussen and Brubaker,5,59 etoxy derivatives were hydrolyzed to 4-methoxyben-
was discussed above (section IIIC). Possible use of zaldehyde. In the case of R = CN, both the α-acetoxy
an artificially low value for the reorganization energy derivative and 4-methoxybenzaldehyde were ob-
of the [CoW12O40]5–/6– couple might have contributed served.
to the high value, 55 kcal mol–1, reported in entry 1. Kinetic studies were carried out spectrophoto-
Because roughly similar results were obtained in only metrically using CoIIIW125– concentrations of 2.5–9.4
two cases (entries 3 and 5, Table 3), reliable deter- × 10–4 M and 100-fold excesses of substrate. A
mination of λ11 for the CoIII/IIW125–/6– couple would mathematical solution of the integrated rate expres-
place Eberson’s conclusions” on more solid footing. sion derived for eqs 57 and 58, in terms of absorbance
In 1993 Baciocchi105 and co-workers studied the as a function of time, was fitted to the experimental
oxidation of a-substituted 4-methoxytoluenes (4- data; values for k1 and the ratio k– 1 /k2 (Table 4) were
MeOPhCH2R; R = H, Me, OH, OMe, OAc, and CN) obtained by nonlinear regression analysis. A plot of
by COIIIW125– in AcOH/H2O (55:45) in order to further log k1 vs ∆G°' values106 for the reactions was fitted
explore the formation and deprotonation of interme- to the Marcus equation (eq 1) using the reorganiza-
diate cation radicals (Scheme 4). tion energy, λ, as an adjustable parameter. A good
fit of the experimental data was obtained for a
Scheme 4 reorganization energy of λ = 41 kcal mol–1.
Because the forward electron-transfer steps (eq 57,
k1) were considerably endergonic, the rate constants
for back electron transfer, k–1, were all in the
diffusion-controlled range and could be assumed
constant. Hence, the data in Table 4 indicated that
as the reduction potentials, E°, of the ArCHR• + /
The mechanism of oxidation of 4-methoxytoluene ArCH2R couples became more positive, the rate of
(p-methoxytoluene, PMT), as described by Eberson,59 forward electron transfer (k 1) decreased, while that
involved two potentially rat&limiting steps (see also of the deprotonation step (k 2) increased. The former
eqs 51 and 52) trend was consistent with the effect of larger positive
∆G°' values on k1, the latter with decreasing stabili-
ties of intermediate cation radicals. However, al-
though cation-radical intermediates derived from
less easily oxidized substrates were more rapidly
deprotonated, k1 values were considerably more
sensitive to changes in E° than were k2 values. As a
result, overall oxidation rates (t1/2 values, Table 4)
followed by rapid formation of products (X = OH or decreased as the reduction potentials of the ArCHR• + /
OAc; see also eqs 54 and 55): ArCH2R couples became more positive.
Finally, using a value of λ = 25 kcal mol–1 for
the reorganization energy associated with the
CoIII/IIW125–/6– electron self-exchange reaction,59,77 a
reorganization energy for the ArCH2R• + /ArCH 2 R
couple of 57 kcal mol–1 was calculated. The authors
noted that although high, the reorganization energy
134 Chemical Reviews, 1998, Vol. 98, No. 1 Weinstock

Scheme 5 Scheme 6

performed a detailed kinetic and product study of the


reactions of 1-arylpropanols, 1-aryl-1,2-propanediols,
and their corresponding methyl ethers, and 1,2-
diarylethanols, with CoIIIW126– in aqueous acetic acid
in the presence and absence of base (AcO–). The
compounds studied possessed the generalized struc-
tures shown in Scheme 5, and the fragmentation of
assigned to the ArCH2R• + ArCH2R couple agreed with intermediate cation radicals was expected to occur
that obtained by Eberson77 (55 kcal mol–1; cross- via either C–H or C–C bond cleavage (pathways a
reaction of p-methoxytoluene with CoIIIW125–). Such or b). The 12-tungstocobalt(III)ate ion was used to
high reorganization energies have been justified on reliably oxidize the parent compounds to cation-
the basis of the strong hyperconjugative effect of the radical intermediates.
α-CHa group, a phenomenon confirmed by ESR Ten compounds of the form shown in Scheme 5
measurements. However, Baciocchi pointed out that were studied in detail, and both Ca–H and Ca-C–b
theoretical calculations did not show any substantial bond cleavage were observed. The reactions of three
differences in molecular geometry between toluene of these compounds–1-(4-methoxyphenyl)-2-meth-
and the toluene radical cation.107 Underestimation oxypropane, 1-phenyl-2-methoxypropane, and 1-(4-
of the reorganization energy of the CoIII/IIW125–/6– methoxyphenyl)-2-propanol–were exemplary (Scheme
couple might also have contributed to the high 6).
reorganization energy calculated for the ArCH2R• +/ Reaction of the α-methoxy derivative 1-(4-methoxy-
ArCH2R couples. phenyl)-2-methoxypropane with CoIIIW125– gave only
An additional manifestation of the kinetic signifi- products of Cα –H bond cleavage (pathway a, Scheme
cance of k1 (electron transfer) and k2 (deprotonation 5). Removal of the 4-methoxy group (1-phenyl-2-
by reaction with base) was demonstrated by Baci- methoxypropane) resulted in a dramatic change in
occhi, Galli, and co-workers.108 They noted that in reactivity: Only products of Cα –Cβ bond cleavage
H-atom-transfer reactions, the order of C-H bond (pathway b) were observed. Oxidation of the β-hy-
selectivities was tertiary > secondary > primary. droxy derivative 1-(4methoxyphenyl)-2-propanol gave
However, in outer-sphere electron-transfer reactions, products of both Cα –H and Cα –Cβ bond cleavage.
compounds possessing secondary C-H bonds were It was reasoned that the 4-methoxy group in the
oxidized most rapidly. ring favored Cα –H bond cleavage by opposing the
Overall reaction rates depend on the values of both accumulation of positive charge on the scissile Cα –
k1 and k2. Moreover, changes in the structures of Cβ bond. In addition, while the presence of both OH
alkyl side chains (e.g., methyl, ethyl, or isopropyl) and OCH3 groups in the alkyl side chain enhanced
present in a homologous series of alkylaromatic com- Cα –Cβ bond cleavage rates, the effect of OH was
pounds have only small effects on oxidation poten- larger than that of OCH3. Kinetic data, including
tials. Thus, in the outer-sphere oxidation of such a solvent isotope effects, suggested that hydrogen
homologous set of compounds, k2 values carry sig- bonding or specific solvation played a key role in the
nificant kinetic weight. In addition, it was argued enhancement of Cα –Cβ bond cleavage by α- or β-OH
that steric and electronic factors favored the abstrac- substituents.
tion of secondary (over primary or tertiary) protons The reactivity of one a-OH-substituted arylpro-
in cation-radical intermediates. When CoIIIW125– panol studied by Baciocchi, 1-(4methoxyphenyl)-2,2-
was reacted with 4-methoxyalkylbenzene derivatives dimethyl-1-propanol (Scheme 7),109 is notable in view
possessing methyl (primary Cα –H bonds), ethyl (sec- of earlier work by Kozhevnikov.110 Kozhevnikov
ondary Cα−Η bonds), and isopropyl (tertiary C α –H studied the oxidation of 1-phenyl-2,2-dimethyl-1-
bond) side chains, the following relative reaction propanol (Scheme 7) by several inorganic oxidants
rates were observed: Et (2.1) > i-Pr (1.8) > Me (1). ([PdCl4]2+, Ce(IV), S2O82–, and [VO(H2O)x ]2+) and by
Baciocchi, Steenken, and co-workers used CoIIIW125– a series of phosphomolybdovanadate complexes
to study Cα –Cβ vs Cα –H bond cleavage and the role ([PV n Mo12–n O40](3+n )–).
of α- and β-OH and -OCH3 groups in the side-chain The difference in product distributions observed for
fragmentation of arylalkanol radical cations.109 They oxidation by Pd(I1) (100% Cα-H bond cleavage) vs
Electron-Transfer Reactions of POMs Chemical Reviews, 1996, Vol. 98, No. 1 135

Scheme 7 Scheme 8

results support Kozhevnikov’s reasoning that the


for oxidation by Ce(IV) (primarily Cα -Cβ bond cleav- products of Cα –H cleavage observed in the oxidation
age) was attributed to the nature of the electron- of 1-phenyl-2,2dimethyl-1-propanol (no 4-methoxy
transfer step. Hydride abstraction by Pd(II) gave group in the ring, Scheme 7) probably arose via
100% Cα –H bond cleavage products, while oxida- mechanistic pathways that did not involve cation-
tion by Ce(IV), a single-electron oxidant, occurred radical intermediates. One reviewer has also pointed
via the formation-by either inner- or outer-sphere out that deprotonation from 1-aryl-2,2-dimethyl-1-
mechanisms-of intermediate arylpropanol radical propanol radical cations–whether a p-methoxy group
cations. Fragmentation of these radicals appeared is present or not-is unlikely due to stereoelectronic
to occur primarily via Cα -Cβ cleavage. effects. Hence, where C–H bond cleavage is ob-
For oxidation by aqueous V(V) and by phospho- served, a different mechanism (probably H-atom
molybdovanadate complexes in water, the following transfer) is probably involved.
Cα –H/Cα -Cβ product distribution ratios were Baciocchi and co-workers also used the CoIIIW125–
observed: [VO(H2O)x ]2+ (0.17), [PVMo11O40]4– (3.9), anion to study the electron-transfer-initiated carbon-
[PV2Mo10O40]5– (3.1), and [PV3Mo9O40]6– (2.0). Be- silicon bond cleavage reactions of ring-substituted
cause the phosphomolybdovanadate complexes are benzyltrialkylsilanes (Scheme 8).107
labile with respect to loss of V(V) ions to solution, In all but one case ((p-methoxybenzyl)triisopropyl-
Kozhevnikov proposed that the Cα –Cβ cleavage prod- silane), the reactions were strictly first order in
ucts resulted from single-electron oxidations by dis- CoIIIW125– and in benzylsilane. Added salts (NaClO4
sociated V(V) ions, while the dominant pathway or NaOAc) decreased the reaction rate, while addition
Cα –H bond cleavage) was indicative of multielectron of CoIIW126– had no retarding effect. By contrast, in
oxidation by the phosphomolybdovanadate com- the oxidation of (p-methoxybenzyl)triisopropylsilane,
plexes. On the basis of this reasoning, however, it the rate was no longer first order in substrate, and
is not obvious how to explain the result obtained added CoIIW126– had a retarding effect. These data
using [PVMo11O40]4–: Either V(V) was reduced by two were interpreted in terms of an electron-transfer
electrons to V(III) or the V(V) ion and a Mo(VI) ion mechanism in which benzyltialkylsilane cation radi-
were each reduced by one electron. cal intermediates underwent Ca –Si bond cleavage to
In oxidations by [PVn Mo 12– n O 4 0 ] ( 3 + n )–, an inner- give benzyl radicals that were further oxidized to
sphere mechanism, involving coordination of the benzyl cations by a second equivalent of CoIIIW125–.
alcohol to a V(V) ion in a fragment of the labile The final step was reaction of the benzyl cations with
vanadium-containing polyanion, was suggested. The solvent (H2O or HOAc). In most cases, initial elec-
formation of vanadate esters, [ROVvO]2+, prior to the tron transfer from the substrate to CoIIIW125– was the
oxidation of alcohols by V(V) was cited as a support- rate-determining step; in the case of (p-methoxyben-
ing precedent.111 Because polyanions can be reduced zyl)triisopropylsilane, cleavage of the Cα –Si bond in
by more than one electron, it was proposed that the the cation-radical intermediate was rate limiting.
reaction proceeded via inner-sphere transfer of two The changeover in mechanism, indicative of a much
electrons from the alcohol to the polyanion. This lower rate of Cα –Si bond cleavage in the (p-meth-
proposal is somewhat puzzling. While 2-e– oxida- oxybenzyl)triisopropylsilane cation radical, was at-
tions of alcohols by high-valent-metal cations are tributed to the presence of the bulky isopropyl groups
well-known, these reactions occur via hydride- on the silicon atom. These groups decreased the rate
transfer mechanisms (the reactions are associated of Cα –Si bond cleavage by stabilizing the silicon atom
with large kinetic isotope effects). Also, the unimo- in the radical cation toward nucleophilic attack by
lecular decomposition of vanadate esters is known to solvent.
occur in a homolytic fashion, resulting in l-e- reduc- On the other hand, orientation of the Cα –Si bond
tion of V(V) to V(IV).111 with respect to the aromatic ring (stereoelectronic
In the oxidation of 1-(4-methoxyphenyl)-2,2-dim- effect) did not appear to influence the rate of C α –Si
ethyl-1-propanol by CoIIIW125– (Scheme 7), Baciocchi’s bond cleavage.112 Oxidation of 2,2-dimethyl-2-silain-
observed 100% Cα -Cβ bond cleavage, despite the dan (Chart 1) by CoIIIW125– was first order in both
presence of the 4-methoxy group, which would have oxidant and substrate, and the value of the bimo-
been expected to suppress the rate of Cα –Cβ cleavage lecular rate constant (1.1 × 10–2 M–1 s–1) was
of the radical-cation intermediate. Thus, Baciocchi’s unaffected by addition of CoIIW126–.
136 Chemical Reviews, 1998, Vol. 98, No.1 Weinstock

Chart 1. Structure of 2,2-Dimethyl-2-silaindane would require further study. To this remark should
be added Eberson’s suggestion (above in this section)
that direct electron exchange between ArCH3• + and
ArCH3 (i.e., oxidation of ArCH3 by ArCH3• + ) might
not occur via an outer-sphere mechanism.77

8. Cyclohexadienes
The Keggin structural class of vanadomolybdophos-
phates are versatile and selective catalysts for ho-
mogeneous O2 oxidations of numerous organic and
Despite the unfavorable stereoelectronic situation inorganic substrates. 113,114 Thus, despite the com-
(the Cα –Si bond could not align itself with the plexity of their solution and redox chemistry, which
aromatic π system), almost 100% yields of Cα –Si make detailed kinetic analysis extremely difficult,
bond-cleavage products were found. No products of efforts to elucidate the mechanisms of the reactions
Cα –H bond cleavage, the stereoelectronically favored particular to these systems114 are to be welcomed (see
pathway, were detected. also sections VIII.A and F).
Bimolecular rate constants, k1, were measured for In 1989 Neumann and Lissel reported the catalytic
the oxidations of ring-substituted benzyltrimethyl- aromatization (dehydrogenation) of cyclic dienes by
silanes by CoIIIW125–.107 By using nonlinear least- the mixed-addenda vanadomolybdophosphate com-
squares regression, the Marcus equation (eq 1) was plex H5[PV2Mo10O40] in the presence of tetraethylene
fit to kinetic (k1) and thermodynamic (∆ Go') data, glycol dimethyl ether (tetraglyme) and O2 in meth-
with reorganization energy, λ, as an adjustable ylene chloride.115
parameter. The best fit gave a reorganization energy
of λ = 43 kcal mol–1. Introduction of the frequently In a more detailed investigation, Neumann and
accepted value of λ for the CoIII/IIW125–/6– electron self- Levin116 provided evidence that stable organic radi-
exchange reaction (25 kcal mol–1 gave a reorganiza- cal-POM association complexes formed during the
tion energy of 61 kcal mol–1 for the ArCH2SiMe3•+ / homogeneous dehydrogenation of α-terpinene to p-
ArCH2SiMe3 couple. This value was smaller than cymene by H5[PV2Mo10O40]•(H2O)n, in acetonitrile (eq
that calculated for the SiEt4• + /SiEt4 couple (70 kcal 61).
mol–1 from the cross-reactions of SiEt4 with Fe(III)
complexes.104 The authors proposed that the larger
value found for SiEt4 was plausible considering the
difference in inner-sphere reorganization energies
associated with removal of an electron from a C–Si
r orbital (SiEt4) vs from the aromatic π system of
ArCH2SiMe3.
The λ value found for the ArCH2SiMe3• + /ArCH 2 -
SiMe3 couple was also larger than those determined
for ArCH3• + /ArCH3 (~40–55 kcal mol–1 by cross-
reaction with CoIIIW125– (55 kcal mol–1) and other One of the difficulties encountered in interpreting
inorganic oxidants.77 This suggested that replace- kinetic data obtained from homogeneous reactions of
ment of a hydrogen atom by a trimethylsilyl group vanadomolybdophosphate anions is that in aqueous
in the a carbon of the side chain influenced the solution, mixtures of heteropolyanions are usually
reorganization energy associated with removal of an present. For example, when dissolved in D2O, samples
electron from the aromatic system. of H5[PV2Mo10O40]•(H2O)n, prepared and isolated ac-
More generally, the relatively higher λ values of cording to the method cited117 by Neumann and
substituted (ArCH3•+ /ArCH3) vs unsubstituted (Ar• + / Levin, give rise to 31P and 51V NMR signals attribut-
Ar) aromatics (~10 kcal mol–1 had been attributed able to a mixture of compounds: H3+n [PV n Mo 12–n O40],
to strong hyperconjugative effects of the CH3 group where n = 1, 2, and 3 and in which the n = 2 species
in the radical cation, which might have led to dominate.86,118–120 Positional isomers of the polyva-
redistribution of C-H bond electron densities.95 nadium anions [PV2Mo10O40l5– and [PV3Mo9O40]6– are
Thus, Baciocchi reasoned, one might expect an even also apparent in NMR spectra.121 The speciation
larger λ value for the ArCH2SiMe3• + /ArCH2SiMe3 chemistry of this labile class of compounds (vanado-
couple because r–π conjugation in a β C–Si bond is molybdophosphates of the Keggin structural type) in
more effective than like conjugation in a β C–H bond. acetonitrile, however, has not been well-character-
However, although ESR measurements have pro- ized. Notably, in 31P NMR spectra of acetonitrile
vided evidence for the hyperconjugative effects of solutions of H5[PV2Mo10O40]•(H2O)n, Neumann116 and
side-chain Cα –H and Cα –Si bonds, ab initio calcula- Levin observed signals corresponding to only one
tions have indicated only very slight changes in bond heteropolyanion, H5[PV2Mo10O40] (mixture of posi-
lengths and angles upon formation of radical cations tional isomers).
from toluene and benzylsilanes. As a result, Bacio- Another complication inherent in the study of
cchi concluded that clarification of the factors re- multielectron oxidations by polyvanadium-containing
sponsible for the reorganization energies associated anions is the capacity of these oxidants to be reduced
with the formation of alkylaromatic radical cations by one or more electrons (reduction of each V(V) ion
Electron-Transfer Reactions of POMs Chemical Reviews, 1998, Vol. 98, No. 1 137

to V(IV)).110,115 In the 2-e– oxidation of α-terpinene,


either 1 or 2 equiv of [PV2Mo10O40]5– might be
reduced.
In an attempt to determine the stoichiometry of the
reaction, increasing amounts of α-terpinene (0–7
equiv) were allowed to react with acetonitrile solu-
tions of H5[PV2Mo10O40]•(H2O)n (8 mM). After each
reaction, the final absorbance of the heteropoly blue
product (750 nm) was recorded. Irreversible reduc-
tion of [PV2Mo10O40]5– by α-terpinene would have
resulted in a linear increase in absorption until a
maximum was reached at the required [substrate]
to [POM] ratio. However, as the number of equiva-
lents of α-terpinene increased, asymptotic behavior,
indicative of a reversible reaction between POM and
substrate, was observed.
Free-energy calculations indicated that the reverse
of 2-e– oxidation of α-terpinene by a single equivalent
of H5[PVv2Mo10O40] (i.e., the reverse of eq 62) was
highly endergonic (∆G° l +44 kcal mol–1. NADH to NAD+, aromatization of the heterocyclic
diene (dihydronicatinamide) ring to the aromatic
cation (N-alkylnicotinamide substituent) occurs via
the loss of two electrons and one proton.122–124
Using electrochemical and spectroscopic tech-
niques, Nadjo demonstrated that in aqueous solu-
This fact forced the authors to contemplate reversible tion at pH 7 (phosphate buffer), 2 equiv of α 2 -
formation of a higher-energy intermediate: [P2VvW17O62]7– were rapidly reduced by each equiva-
lent of NADH:

The nature of the oxidation product and analogy of


eq 65 to biological oxidation were confirmed by bio-
The suspected intermediate was observed by ESR chemical reduction (addition of alcohol dehydroge-
spectroscopy. The intermediate was stable for sev- nase and ethanol) of the product, NAD+, back to
eral hours at room temperature: ESR spectra, which NADH.
confirmed that the POM anion was reduced by only In an elegant study, Nadjo used both stopped-
one electron, remained unchanged after 6 h. Support flow123 and electrochemical (double potential step
for the formation of an organic radical was provided chronocoulometry, DPSC)124 methods to explore the
by the observation of new lines in the ESR spectrum kinetics of oxidation of NADH by a series of mixed-
upon addition of the spin trap 2-methyl-2-nitroso- addenda Wells-Dawson anions. The reduction po-
propane. tentials of the heteropolyanions studied spanned a
Possible mechanisms for conversion of the radical range of ~250 mV: α2-[P2MoVIW17O62]6– (E° = +466
intermediate to p-cymene were discussed. The ab- mV vs NHE), [P2V2VW16O62]8– (+519 mV), α2 -
sence of oxygenation products observed after catalytic [P2VvW17O62]7– (+634 mV), α7– VI 6–
1-[P2Mo W17O62] (+637
oxidation of α-terpinene by H5[PVv2Mo10O40] in the V
mV), and α1-[P2V W17O62] (+719 mV). In both
presence of O2 argued against diffusion of organic studies, l-e- reduction of the Wells-Dawson anions
radical intermediates into solution. Although no was observed and the stoichiometry shown in eq 66
conclusion was reached, the authors considered fur- obeyed. The reactions were first order in POM and
ther unimolecular decay of the association complex in NADH.
(electron transfer followed rapidly by proton transfer For the set of reactions studied, the straight line
or perhaps H-atom transfer; eq 64) an attractive (R = 0.999 987) obtained by plotting log k (k =
possibility. bimolecular rate constant) vs E° (reduction potential
of each POM anion) had a slope of +16.4 V–1 (Figure
5; see section II.D).
In theory, for a series of endergonic electron-
transfer reactions in which electron transfer within
encounter complexes is reversible and where the
Keita, Nadjo, and co-workers studied the 2-e– rates of back-electron transfer compete effectively
oxidation of dihydronicatinamide adenosine dinucle- with dissociation of the successor complexes, the plot
otide (NADH) to nicatinamide adenosine dinucleotide of log k vs E° (E° = anodic potential or reduction
(NAD) and H+ by Wells-Dawson heteropolytung- potential of an oxidant) should give a straight line
states (POMox, eq 65). In the biological oxidation of with a slope equal to +16.6 V–1 (see eq 34 in section
138 Chemical Reviews, 1998, Vol. 98, No. 1 Weinstock

Table 5. Relative Rates of Oxidation of Substituted


Arylacetic Acids by CoIIIW 125– and NiIVMo 96– in Acetic
Acid/KOAc

occurs at a substantially more positive potential than


those associated with the reduction of CoIIW126– to
Figure 5. Plot of log k vs E° (reduction potential vs NHE [CoIIW12O40]7– or [CoIIW12O40]8–.52,88,99 In addition,
of each POM anion) for the oxidation of NADH by mixed- good Hammett correlations were found for all but the
addenda (V(V))- and Mo(VI)-substituted) Wells-Dawson methoxy-substituted compounds, which reacted more
tungstophosphate anions. rapidly. For oxidations by NiIVMo96–, and excluding
the methoxy-substituted derivatives, a Hammett ρ
II.D). As described in section II.D, the bimolecular value of –1.71 was calculated, and for CoIIIW125–, a
rate constant, k, can be described as the product k = ρ of –0.77. Both values were reported to be in range
K 1 K e tk3, where K1 = the equilibrium constant for for a process involving rate-limiting formation of a
formation of the encounter complex, Ket = equilibri- benzyl radical (ρ = –1.5 to –0.31.
um constant for electron transfer within the encoun- In addition, good correlations (log-log plots) of the
ter complex, and k3 is the rate of dissociation of the rates of oxidation by NiIVMo96– (again excluding the
successor complex. Therefore, the observed slope methoxy-substituted derivatives) with the relative
(16.4 V–1) provided strong evidence that separation rates of decarboxylation of the same set of substrates
of the radical ion pair was the rate-determining step; by three other oxidants–CoIII(OAc)3 Cu(III), and Ce-
that the reaction was thus under thermodynamic (IV)–suggested similar mechanisms. In oxidations
control (i.e., dictated by Ket, which is equal to kfet/ by Ce(IV), rate-limiting decomposition of metalloor-
kret), and that oxidation step (k fet) involved 1-e– ganic intermediates [ArCH2CO2Ce2IV]3+ were pro-
oxidation of NADH to the radical cation NADH•+ . By posed.125 Jönsson, however, on the assumption that
extrapolation to an estimated diffision-rate-limited CoIIIW125– and NiIVMo96– acted exclusively as outer-
value of ln k (from the plot of log k vs E°, Figure 5), sphere oxidants, proposed that both POM-oxidation
the potential of the NADH• + /NADH couple was calcu- reactions proceeded via outer-sphere oxidation of the
lated. Close agreement between their experimental carboxylic acid group (eqs 68–71).
value, E° = 1.05 V (NHE), and one literature value
(1.00 V) provided satisfying closure to this study.

C. Organic Acids, Esters, Aldehydes, and

The oxidation of arylacetic acids by CoIIIW125– and


[NiIVMo9O32]6– (NiIVMo96–) was reported by Jönsson
in 1983.98 The reaction between CoIIIW125– or NiIVMo96–
and arylacetic acids in refluxing acetic acid, in the
presence of 0.5 M KOAc and acetic anhydride,
resulted in the formation of benzyl acetates in greater
than 80% yields: This mechanism is analogous to that seen in the
anodic oxidation of carboxylic acids (Kolbe mecha-
nism): Electron transfer from the carboxylic acid
group is followed, after loss of CO2, by oxidation of
the benzylic radical.
The assumption that oxidation by NiIVMo96– oc-
Arylacetic acids studied and relative rates of oxida- curred via an outer-sphere mechanism may have
tion by CoIIIW125– and NiIVMo96– are shown in Table been incorrect. In related work,126 Jönsson observed
5. very similar behavior to that of NiIVMo96– (ρ = –1.71)
log-log plot of the rate data in Table 5 gave a when a series of arylacetic acids were oxidized by Cu-
linear correlation of 0.97, indicative of similar rate- (III) in refluxing acetic acid (a ρ value of –1.4 for
limiting steps when oxidized by either anion. This arylacetic acids with less electron-donating substit-
was significant considering that NiIVMo96– is rapidly uents and unusually rapid rates of oxidation of
and irreversibly reduced by two electrons, while methoxy-substituted derivatives). An inner-sphere
multiple-electron reductions of CoIIIW125– are revers- mechanism involving intermediate [ArCH2CO2CuIII]2+
ible, and l-e- reduction of CoIIIW125– to CoIIW126– complexes was proposed. The results obtained using
Electron-Transfer Reactions of POMs Chemical Reviews, 1998, Vol. 98, No. 1 139

NiIVMo96– did not clearly exclude rate-limiting de- transition states of the reaction. The results dem-
composition of related inner-sphere complexes. onstrated that the rate of the outer-sphere electron-
The unusually rapid rate of oxidation of the ρ− transfer step (eq 75) was highly sensitive to low
methoxy-substituted derivative was attributed to an concentrations of added solutes (in this case alcohols)
alternative mechanism: direct oxidation of the elec- capable of forming 1:1 encounter complexes. In
tron-rich aromatic moiety to a radical cation, followed addition, greater rate retardation was observed as
by loss of carbon dioxide and a proton, giving a the number of methylene groups in the alkyl side
benzylic radical that was further oxidized to the chain of the added alcohol increased. This was
cation (eqs 72–74). interpreted as indicating a loss of hydrophobicity in
the transition state, relative to that of the initial
state, of the reaction,
Banejee and colleagues, in a carefully executed
study, provided clear evidence for the formation of
stable ternary (donor–M+–acceptor, M = alkali-
metal cation) complexes prior to electron transfer in
the oxidative decarboxylation of aliphatic organic
acids by aqueous CoIIIW125– in solutions of very high
ionic strengths (µ = 1-2 M added electrolyte).
(Electron-exchange and electron-transfer reactions of
heteropolyanions were reviewed by Saha, Ali, and
Banerjee in 1993.)129 Oxidations of oxalate,130 for-
Trahanovsky125 had previously attributed unusu- mate,131 citric acid,132 and a variety of amine-N-
ally rapid rates of oxidation of methoxy-substituted carboxylates (iminodiacetate, nitrilotriacetate, and
arylacetic acids by Ce(IV) to a like mechanism.126 ethylenediaminetetraacetic acid, H4edta)133 and his-
In 1996, Bietti, Baciocchi, and Engberts returned tidine134 were studied. In each case, due attention
to the oxidation of ρ-methoxyphenyl acetate by was paid to the effects of pH, ionic strength, and
CoIIIW125– to study the effects of solvent composition added alkali-metal cations (K+, Na+, and Li+) or
on reaction rate.127 On the basis of a theoretical anions (ClO4–, SO4–, and NO3–). Indicative of the
model developed by Engberts for the quantitative care exercised in these studies was the measurement
analysis of medium effects on reaction rates in highly of rate constants for both uncatalyzed and alkali-
aqueous binary solvents, 128 the rate of oxidation of metal cation catalyzed oxidations of H4edta as a
the ρ-methoxyphenyl acetate anion (KH2PO4/KNaH- function of pH (i.e., rate constants for oxidations of
PO4 buffer at pH 7.4 in water) was measured as a H4edta, H3edta–, H2edta2–, and Hedta3–).133 Progress
function of added alcohol (0–3 mol % ROH; R = Me, of reaction was measured spectrophotometrically by
Et, n-Pr, n-Bu, t-Bu, or CF3H2) concentration. observation of Co(III) (λ max = 388 nm) and/or Co(II)
Kinetic data were consistent with the mechanism (λ max = 625 nm), and 2:1 reaction stoichiometries (i.e.,
shown in eqs 75–77 and even at low concentrations, 2CoIIIW125– per equivalent of organic acid) were
all of the alcohols induced large decreases in reaction observed in each case (cf., oxidation of oxalate, eq 78).
rates. For example, the rate constant in 1.5 molal
(m) 1-propanol (91.5 g of 1-propanol per kilogram of
solvent mixture or ~3 mol % ROH) was ~27% of that
in water. Well-defined isosbestic points in the time-depend-
ent visible spectra of reaction mixtures argued against
the accumulation of appreciable amounts of reaction
intermediates. The intermediacy of organic radicals
in the multistep reactions was demonstrated by the
polymerization of acrylonitrile, which occurred only
when all components of the reaction were present.
Reactions were carried out under pseudo-first-order
conditions using submillimolar concentrations of
CoIIIW125– and 100- to 500-fold excesses of organic
acids.
The oxidation of oxalate was studied at ~1 M ionic
strength. First-order dependence of reaction rates
on CoIIIW125– and oxalate concentrations (pH = 5)
suggested a rate law of the form

As predicted, plots of –ln(km /kw) vs m of added


alcohol (km = rate constant in the presence of m molal
alcohol; kw = rate constant in buffered water alone)
were linear. According to Engbert’s model, the slope
of each line was proportional to the difference be-
tween the pairwise Gibbs energies associated with
the 1:1 interaction of added alcohol with initial vs
140 Chemical Reviews, 1998, Vol. 98, No. 1 Weinstock

Measurement of reaction rate, –d[CoIIIW125–]/dt, as and probably held together by electrostatic forces,
a function of added K+ revealed a linear dependence formed prior to electron transfer.
of k1 on [K+]: Alternatively, the elementary steps leading to
reduction of the first equivalent of Co(III) might have
been depicted more explicitly in terms of an equilib-
rium constant for ternary complex formation, Kc o m p l e x
At 30 °C, values of k0 = 12.4 × 10–4 M–1 s–1 and k' = (M–2), and a unimolecular rate constant for intramo-
10.2 × 10–3 M–2 s–1 were obtained. These values lecular electron transfer, ket (s–1) (waters of hydration
were typical (within a few orders of magnitude) of have been ignored):
those found for catalyzed and uncatalyzed oxidations
of a variety of acids and their conjugate bases.130–134
Similar results were obtained for HC2O4– at pH 1.32.
Catalysis by alkali-metal cations followed the order
K+ > Na+ > Li+, while neither the nature of the
anions present (ClO4–, SO4–, or NO3–) nor changes
in ionic strength at constant alkali-metal cation
concentrations, had significant effects on reaction
rates. In catalysis of the oxidation of HC2O4– by Na+,
a second-order dependence on [Na+] was observed: Spontaneous self-assembly of the ternary complex
anion, A, which exists in solution at concentrations
dictated by the formation constant, Kcomplex, is fol-
lowed by intramolecular electron-transfer associated
with the unimolecular rate constant, ket (see section
II.D). First-order dependence of absorbance on
Inverse relationships between reaction rates and [H+] [CoIIIW125–] suggests that dissociation of the Co(II)
were quantitatively correlated with the pKa values complex, B, occurred prior to rapid oxidation of the
for acid dissociation of oxalic acid and qualitatively oxalate radical anion, C2O4• –, by a second equivalent
rationalized in term of the effect of protonation state of CoIIIW125–. (This was not the case in the Na+-
on the oxidation potential of C2O42–. catalyzed oxidation of formate,131 in which second-
A lack of dependence of reaction rate on ionic order dependence of absorbance on the concentration
strength ruled out an effect of ionic strength on of COIIIW125– suggested that formate-radical anions
Coulombic work terms (eqs 1 and 36) as a source of in complexes analogous to B might have been oxi-
the rate enhancement. Moreover, the strictly linear dized by a second equivalent of CoIIIW125– prior to
or second-order dependencies of reaction rates on dissociation into component species.) The rate ex-
alkali-metal cation concentrations suggested that pression for oxidation of oxalate by CoIIIW125– might
ternary complex formation (section II.D), rather than thus have been expressed as the product:
ion pairing alone (i.e., [(M+)(POM n –)](n –1)– M =
alkali metal; section II.C), was responsible for the
observed rate enhancements. Additional evidence
for ternary complex formation was provided by
variable temperature studies of the uncatalyzed and
Na+-catalyzed oxidations of formate anion.131 Reduc-
tions in ∆H‡ values associated with Na+ catalysis Although this expression has the same form as those
(from ∆H‡ = 35 to ∆H‡ = 24 kcal mol–1) were too used by Sutin to describe outer-sphere electron-
large to be accounted for by M+–POM ion pairing transfer processes (section II.B.2, eq 10), the equi-
alone (i.e., by changes in effective charge products, librium constant, Kcomplex, in eq 87 refers to the
z1 z2 (eq 1)). formation of a stable ternary complex (section II.D).
In the case of Na+-catalysis of oxalate (C2O42–) Linear or second-order dependences of rates on
oxidation, the following reaction mechanism was alkali-metal cation concentrations (K+ > Na+ > Li+)
suggested: were also observed in the oxidation of citrate,132
iminodiacetate, nitrilotriacetate, ethylenediamine-
tetraacetic acid (H4edta),133,135,136 N-(2-hydroxyethyl)-
ethylenediamine triacetate,137 α-hydroxyacids (man-
delic, lactic, and glycolic),138 and histidine134 at high
ionic strengths. In the 2-e– oxidation of citric acid
at high ionic strengths (eq 88), limiting dependences

An activated complex of the form “[CoIIIW12O40]5–•••


Na+•••C2O42–]”, in which the Na+ cation acted as a
bridge, was proposed for the rate-limiting step (eq of rates on citrate anion (H2A–, HA2–, and A3–)
83). More conservatively, the kinetic data suggested concentrations at very high citrate anion to CoIIIW125–
that ternary complexes, of indeterminate structure ratios (i.e., saturation kinetics) provided additional
Electron-Transfer Reactions of POMs Chemical Reviews, 1998, Vol. 98, No. 1 141

evidence for the formation of donor-acceptor com-


plexes prior to electron transfer.
In the oxidation of cyclohexanone, 2-butanone, and
1,3-dihydroxy-2-propanone by CoIIIW125– at high ionic
strengths (0.5–1.0 M Na+), the reaction was first- The reaction was first order in [CoIIIW125–] over a
order in oxidant at high [H+] (0.30–0.40 M HClO4) range of pH values, throughout which an inverse
and low [CoIIIW125–] (0.1–0.2 mM) but zeroth-order dependence on [H+] was observed. Specific alkali-
in oxidant at low [H+] (0.02–0.0006 M) and high metal cation catalysis was observed as well, but in
[CoIIIW125–] (5.0 mM).139 These data were consistent an unexpected order: Li+ > Na+ > K+. For the Na+-
with a mechanism in which acid-catalyzed enoliza- catalyzed reaction, the following elementary steps
tion occurred prior to oxidation by CoIIIW125– (eqs 89 and rate expression were proposed (RH = ethyl
and 90; R• is an enolyl radical). ester):

Assuming –d[enol]/dt = 0 (steady state in [enol])


and that the enolyl radical formed in eq 90 was
oxidized rapidly by a second equivalent of CoIIIW125–, Inverse dependence on [H+] was tentatively at-
the following rate expression was derived: tributed to stabilization of the free radicals, R• (Chart
2). It was proposed that stabilization of these
intermediates increased the rate of the reverse path
of the rate-determining step (eq 95).
Chart 2. H+-Stabilized Ethyl Ester Radical
Intermediates
Under conditions at which enolization was rate
limiting (zeroth order in [CoIIIW125–]), reaction rates
were independent of [Na+]. Catalysis by alkali-metal
cations was not investigated under conditions at
which reaction rates exhibited a pseudo-first-order
dependence on [CoIIIW125–].
Unable to correlate reaction rates with those The unusual order of the alkali-metal cation de-
predicted by Marcus theory, it was proposed that pendence was tentatively attributed to ion-induced
“since the central [Co] ions are surrounded by WO4 dipole interactions exerted by the alkali-metal cation
octahedra rather than by oxygen atoms, the transfer on the neutral ester (RH in the ternary complex, C,
of electrons across the octahedra may not be equiva- eq 95). Thus, the poorer the cation at inducing
lent to transfer in simple oxo-anions”. This proposal positive charge in the neutral ester (Li+ < Na+ < K+),
was originally set forth by Rasmussen and Brubaker the more easily and rapidly the neutral ester was
to account for lower-than-theoretical rates of electron oxidized.
self-exchange between CoIIIW125– and CoIIW126– (see However, the order of tetraalkylammonium and
section III.C).5 alkali-metal cation catalysis in the reduction of
However, in calculating the thermodynamic driving organic substrates by CoIIW126– is opposite141 to that
force (∆G° or E°) for the reaction, Banerjee and observed in oxidations by CoIIIW125–. Hence, if the
coauthors chose a value for the reduction potential oxidation of RH to [R•H]+ was in fact reversible, the
of CoIIIW125– (1.0 V vs NHE) that was not applicable52 observed order of alkali-metal cation catalysis might
at the ionic strengths used. More generally, failure simply have reflected the effect of ion pairing on the
to observe correlations between Marcus theory and rate of reduction of [R•H]+ by [CoIIW12O40]6– in D (k – 2
rates of oxidation of ketones by CoIIIW125– were likely and the ratio k2 /k–2, eq 96; eq 97 is added for
due to ion pairing, complex formation, and related completeness, although other elementary steps are
phenomena (see sections II.C and D), rather than to possible).
a unique property of the [W12O40]8– ligand.
In the oxidation of simple esters (ethyl acetoace-
tate, eq 92; diethyl malonate, eq 93) by CoIIIW125–
under acidic conditions at high ionic strengths, no
evidence was found for rate-limiting keto-enol tau-
tomerization. 140

Catalysis by alkali-metal cations, rate-limiting


acid-catalyzed enolization, and second-order depen-
142 Chemical Reviews, 1998, Vol. 98, No. 1 Weinstock

the opportunity and facilities needed to obtain further


information, such as would be required to better
define the extent of ion pairing, the nature of the
donor-acceptor complexes, the rate of intramolecular
electron transfer within these and related complexes,
and the possible role of atom (H• ) transfer in rate-
determining oxidation steps (kinetic isotope ef-
fects).143
D. Aromatic and Aliphatic Alcohols
The large and growing number of examples of
homogeneous oxidations of phenols and related com-
pounds by vanadium-containing mixed-addenda het-
eropolymolybdate anions, mostly of the Keggin-
structural type (H3+n [PVn Mo 12–n O40], n = 1–6) have
been adequately reviewed.113,114,144,145
Limited evidence for both outer-sphere146 and
multiple-electron-transfer mechanisms114,145 has been
reported. In the oxidation of a series of alkylated
phenols by H5[PV2Mo10O40], Lissel et al.146 observed
a strong dependence of percent conversions on E°
values of the phenolic substrates. On this basis, an
electron-transfer mechanism was proposed. In the
oxidation of phenols by mixed-addenda vanadomo-
lybdophosphate complexes, H3+n [PVnMo 12–n O40], prod-
dence of reaction rates on [CoIIIW125–] were observed ucts of 1- and 2-e– oxidations were observed and
in the oxidation of D-glucose (Glu), D-mannose (Man), complex mechanisms, involving dissociation of V(V)
and D-fructose (Fru) (Chart 3).142 ions, proposed.147–149
Reactions were carried out under pseudo-first-order In 1977 Amjad, Brodovitch, and McAuley76 studied
conditions ([CoIIIW125–] = 0.2 mM and [sugar] = 800 the oxidation of L-ascorbic acid, hydroquinone, and
mM) at low pH values ([H+] = 0.05–1.00 M, HClO4 catechol (0.7–1.5 mM) by CoIIIW125– (0.1–0.45 mM)
or HNO3) and at high ionic strengths (µ = ~1.0 M, at low pH values ([HClO4] = 0.04–1.0 M) and high
NaClO4). At 60 °C, 2 equiv of CoIIIW125– were reduced ionic strength (µ = 1.0 M, LiClO4] using a stopped-
by each equivalent of sugar, and the oxidations of Glu flow apparatus. Under pseudo-first-order conditions,
and Man gave arabinose (Chart 4). The oxidation of the reactions were first order in [CoIIIW125–] and in
Fru gave a mixture of products, including arabinose. [alcohol], and reaction stoicbiometries of 2:l CoIIIW125–
Kinetic data were summarized by the following to alcohol suggested the formation of quinonoid
experimental rate laws (M = alkali-metal cation): products. For hydroquinone and catechol, no [H+]
dependence was observed over the pH range used,
and the proposed mechanism involved rate-limiting
oxidation followed by loss of a proton:

For Glu, n was equal to 1 for K+ and 2 for Na+ and


Li+; for Man, n was equal to 1 for K+ and Na+ and 2
for Li+. Kate constants for the uncatalyzed reactions
ranged from ~10–1 to 101 (variable units) with values In the oxidation of L-ascorbic acid (H2A), an inverse
of k' in the same range and generally somewhat dependence of rate on [H+] gave an experimental rate
smaller than corresponding values of k. With char- expression that included the parallel oxidation of
acteristic attention to detail, the authors noted that both diprotic (H2A) and monoprotic (HA–) forms of
the rate of mutarotation between α and β anomeric the acid:
forms of Glu was fast relative to the rate of oxidation
by CoIIIW125–.
The rate laws determined by Banejee suggested
extensive ion pairing and the formation of donor-
acceptor complexes, the latter involving both unoxi-
dized sugars and radical intermediates. Given the
care and attention to detail obvious in Banerjee’s
work, his evidence for the presence of these phenom- Comparison of experimental rate constants with
ena must be regarded as highly reliable. It is hoped those predicted by Marcus gave only qualitative
that he and his colleagues might be provided with agreement. (A value of λ for the CoIIIW125–/CoIIW126–
Electron-Transfer Reactions of POMs Chemical Reviews, 1998, Vol. 98, No. 1 143

self-exchange reaction of 61.4 kcal mol–1 was used. (>3 V), outer-sphere oxidation to the corresponding
This value was obtained by extrapolating to µ = 1.0 alkoxy radical (eq 105) is unlikely to occur.
M, using data published by Rasmussen and Brubak- The oxidation of pentane-1,5-diol exhibited a first-
er;5 see section III.C.) More fundamentally, kinetic order dependence on [H+] and second-order depen-
isotope effects, which might have clarified the nature dence on [HOCH2(CH2)3CH2OH]. A set of elementary
of the e–/H+-transfer step, were not performed. In steps, which included donor-acceptor complex for-
addition, neither the dependence of reaction rate on mation and which was consistent with the observed
ionic strength nor on the nature of the alkali-metal participation of water in the rate-limiting step, was
cation present were investigated. Hence, the pro- proposed. Appropriate approximations gave a rate
posed mechanism (eqs 100 and 101), while consistent expression consistent with the kinetic data:
with the experimental rate law, may be overly simple.
The oxidation of aliphatic alcohols (butane-1,3-diol,
butane-1,4diol, and pentane-1,5-diol; 0.3–1.5 M in
each) by CoIIIW125– (0.0002 M, pseudo-first-order in
diol) at very high ionic strength (µ = 2.0 M, NaClO4)
was investigated by Olatunji and Ayoko over a [H+]
range of from 0.75 to 1.75 M, using an initial rate The lack of reactivity of the butane-1,3-diol and
method.150 Reaction of butane-1,3-diol was immea- mechanistic divergence between the 1,4- and 1,5-
surably slow. The 1,4- and 1,5-diols, however, were diols, were rationalized by a suggestion that the “1,3
oxidized by two electrons to 4- and 5-hydroxy alde- and 1,4 diols may...involve a chelate type complex in
hydes (eq 103; n = 2 or 3): the transition state in which the two hydroxyl groups
are closely associated with the oxidant through one
of the tungsten atoms” and that such a transition
state was less likely for the larger 1,5-diol. Consider-
ing the kinetic stability of the tungsten-oxygen
bonds in heteropolytungstates and the lack of an
available coordination site on octahedral W(VI),
The rate of oxidation of butane-1,4-diol was inde- hydrogen bonding between two hydroxyl hydrogen of
pendent of [H+] and first order in both oxidant and the diol and two µ-2 W–O–W oxygens of [CoW12O40]5–
diol. The proposed mechanism included equilibrium is a more reasonable proposal.
formation of a donor-acceptor complex, followed by Banerjee129 has pointed out that the specifics of
reversible, rate-limiting electron transfer: Olatunji’s proposal were unsupported by experimen-
tal data. Nonetheless, the experimental rate law
determined for the oxidation of pentane-1,5-diol did
suggest a complex, multistep reaction mechanism
that may have differed in some fundamental way
from that operative in the oxidation of butane-1,4
diol. Comparison of kinetic results obtained in the
oxidation of the 1,3-, 1,4, and 1,5-diols suggested that
the bifunctional nature of these substrates had
mechanistic consequences.

E. Carbon-Centered Organic Radicals


In the (overall) 2-e– oxidation of organic substrates
(subH) by 2 equiv of oxidized POM anions (POMox,
eqs 109 and 110) the second electron-transfer step
(oxidation of radical intermediates, eq 110) is usually
Assuming a steady-state concentration of the donor- assumed to occur very rapidly.
acceptor complex (eq 104) and that k2[Co IIIW125–] p
k –1[CoIIIW125–] [H+], the rate expression derived from
the proposed mechanism gave the experimentally
observed rate law

In a series of detailed studies involving POM


oxidations of a variety of organic radicals generated
by continuous and pulse radiolysis, Papaconstan-
tinou, 151-155 working in Athens, Greece, single-hand-
Although the proposed mechanism (eqs 104–106) edly proved this assumption correct.
reasonably suggested complexity beyond that needed In an initial study, Papaconstantinou investigated
to account for the experimentally observed rate law, the rate of reduction of aqueous Keggin and Wells-
no evidence for ion-pair formation or for association Dawson heteropolytungstate and molybdate anions
(donor-acceptor) complex formation was provided. In ([PW12O40]3–, [P2W18O62]6–, and [P2Mo18O62]6–) under
addition, one reviewer has pointed out that, because conditions of continuous exposure to 60Co λ - radia-
aliphatic alcohols have very high oxidation potentials 151
tion. The irradiation reactions were carried out
144 Chemical Reviews, 1998, Vol. 98, No. 1 Weinstock

Table 6. Second-Order Rate Constants for the 1-e– Reductions of Heteropolytungstate Anions by Hydroxyalkyl
and Carboxyl Radicals

in the presence of O2, Ar, N2O, several alcohols, and the near-diffusion-controlled rates and 100% efficien-
formic acid and over a pH range of ~1–4. cies of the electron-transfer reactions leading from
The radiolysis of water gives a number of products radiolysis to POM reduction.
(eq 111); numbers in parentheses (G values) indicate The potentials associated with successive stepwise
the number of species, of the indicated type, produced reductions of many POM anions are easily measured
for each 100 eV of radiation energy absorbed by by electrochemical methods. Making use of this
water. thermodynamic data, Papaconstantinou used het-
eropolytungstates and heteropolymolybdates to de-
termine the apparent reduction potentials of several
radiolytically generated hydroxyalkyl and carboxyl
radicals.l152 For each type of radical generated by
In solutions deaerated with Ar at pH ≈ 1, solvated continuous 60Co-λ-irradiation, reduction of the het-
electrons, e–(H2O)n , react rapidly with H+ ions to give eropoly blue continued until the potential of its next
hydrogen atoms, H• . At pH values of ~3–4 in reduction step was more negative than the reduction
solutions deaerated using N2O, the aqueous-solvated potential of the radical. Hence, the reduction poten-
electrons are rapidly consumed in the stoichiometric tial of each radical was estimated by measuring the
production of HO• radicals: degree-of-reduction of the heteropoly blue formed in
its presence.
To obtain kinetic data pertaining to the near-
diffusion-controlled rate constants for reduction of
In the presence of excess organic alcohols or formic POM anions by alkyl radicals, a rapid pulse-radioly-
acid-MeOH, EtOH, n-PrOH, i-PrOH, or HCOOH sis technique was required. With the use of 30 ns
(RH, eqs 113 and 114)–either set of reaction condi- pulses of a 2.3 MeV electron beam to irradiate
tions gives the corresponding carbon-centered radi- aqueous solutions containing POM anions and simple
cals (R•, where R = CH2OH, etc.).151 However, alcohols, formic acid, or formate, second-order rate
constants were determined from oscilloscope traces
of the absorbance-formation kinetics of the 1-e--
reduced POM anions (Table 6).156
Although the rates of most of the reactions were
substantially diffusion controlled, clear trends were
oxidation of RH by HO• (eq 114, k = 108–109 M–1 s–1 evident: For each POM anion, rates of reduction by
is -100 times more rapid than oxidation by H• (eq the hydroxyalkyl radicals increased in the order of
113). their electron-donating abilities, •CH2OH < CH3C•-
Plots of absorbance of the reduced (heteropoly blue) HOH < (CH3)2C•OH; rates of reaction decreased as
anions vs time, from spectroscopic data obtained the reduction potentials of the POM anions moved
during conditions of continuous 60Co-λ -irradiation of to more negative (less thermodynamically favorable)
aqueous solutions of POM anions and alcohols or values.
formic acid (15 POM:RH combinations in total), were Additional factors were evident in the carboxyl
consistently linear. While the tungstophosphates radical reactions.155 As in reduction of the POM
were each reduced by one electron, only 2-e–-reduced anions by the hydroxyalkyl radicals, reaction rates
[P2Mo18O62]8– was observed. This result was at- decreased as the reduction potentials of the POM
tributed to rapid disproportionation of 1-e–-reduced anions became more negative, and reduction of
[P2Mo18O62]7– anions (see below). On the basis of [SiW12O40]4– by •CO2H was slower than reduction by
dosimetric analysis; the sum of the G values associ- the more electron donating alkyl radical, •CO2–.
ated with the radiolytic formation of e–, H•, and HO•; However, as the negative charge of the POMox anion
and the extinction coefficients of the heteropoly blue increased, the rates of 1-e– reduction by •CO2–
species, it was calculated that the reactions (from dropped more rapidly than did those observed for 1-e–
radiolysis to POM reduction) were 100% efficient. The reduction by •CO2H. In fact, when generated in the
zero-order kinetics (linear plots of absorbance, mea- presence of the highly negatively charged anions
sured at 2-min intervals, vs time) were attributed to [FeW12O40]5– and [H2W12O40]6–, the more electron-
Electron-Transfer Reactions of POMs Chemical Reviews, 1998, Vol. 98, No. 1 145

donating radical •CO2– reacted more slowly than did s–. In addition, the rate-retarding effect of the
•CO2H. This effect, attributed to repulsion between negative charge of the carboxy1 radical (•CO2–) was
the negatively charged radical, •CO2–, and the POM noted.
anions, was consistent with the likely outer-sphere However, very different results were observed in
nature of these electron-transfer reactions. the reduction of CoIIIW125– by increasingly branched
Whereas 1-e– reductions of [P2W18O62]– and other acyclic alkyl radicals. In this case, rates of reduction
tungstates are well resolved by electrochemical meth- covered 4 orders of magnitude, from less than 105 M–1
ods, only 2-e–reduction steps are observed for s–1 to the diffusion-controlled limit: CH3• (< 105 M–1
[P2Mo18O62]6–. Using pulse radiolysis, Papaconstan- s–1) ^ CH3CH2• (2.0 × 106 M–1 s–1) ^ (CH3)2CH• (1.3
tinou157 determined the bimolecular rate constant × 109 M–1 s–1) < (CH3)3C• (3.3 × 109 M–1 s–1, probably
(3.9 × 109 M–1 s–1) for 1-e– reduction of [P2Mo18O62]6– diffusion limited). The Taft plot (Hammett linear
to H[P2Mo18O62]6– (pH 2) by (CH3)2C•OH radicals and free-energy relationship extended to aliphatic sys-
recorded the absorbance spectrum of H[P2Mo18O62]6–. tems) gave a slope (ρ* value) of –28, indicative of
From subsequent changes in the absorption spectrum high sensitivity of the reaction rate to the nature of
(the H[P2Mo18O62]6– anion was stable Only in the the alkyl substituent at Cα. By contrast, the Ham-
subsecond time frame), the bimolecular rate constant mett ρ* value measured for reaction of the same set
for disproportionation (1.0 × 104 M–1 s–1) was mea- of radicals with FeIII(CN)6]3– was –13, indicative of
sured: a substantially different mechanism. The Fe(III)
complex, it was pointed out, could react via ligand
transfer to give alkylnitrile products (for example,
reaction of FeIII(CN)6]3– with CH3• gives CH3CN in
20% yield). In addition, Baciocchi found that the
rates of reaction of the same set of alkyl radicals with
Building on this work, Papaconstantinou used the [IrIVC16]2– were all near the diffusion-controlled limit
[P2Mo18O62]6– anion to investigate the redox chem- ((1.2–3.8) × 109 M–1 s–1) and noted that reaction of
istry of fat-soluble α-tocopherol (vitamin E)158 and [IrIVCl6]2– with CH3• gave CH3Cl in 100% yield.
water-soluble ascorbic acid (vitamin C)159 in micellar Taken together, these data indicated that oxidation
media (aqueous aggregates of nonionic surfactants) of the alkyl radicals by CoIIIW125– occurred via an
designed to model biological membranes. Relative electron-transfer mechanism.
to rates observed in isotropic media, the rate of
reduction of [P2Mo18O62]6– by vitamin E decreased by F. Thiols and Organic Sulfides
3 orders of magnitude, while the rate of polyanion Ayoko and Olatunji investigated the oxidative
reduction by vitamin C decreased by only a factor of dimerization of biologically relevant thiols (glu-
10. tathione,162 L–cysteine, mercaptoacetic acid, and β-mer-
163
After Papaconstantinou’s studies, Lerat et al. evalu- captoethylamine (2.0–40 mM)) by CoIIIW125– (0.2
ated the rate constants for reduction of metatung- mM) at high ionic strength (µ = 1.0 M, NaClO4) over
state, [H2W12O40]6–, by hydrated electrons, by H a range of [H+] (0–1.0 M). Although direct evidence
atoms, and by CH3C•HOH radicals generated by for the proposed oxidation products was not obtained,
pulse radiolysis.160 Notably, the rate of reduction of reaction stoichiometries were consistent with the
metatungstate by hydrated electrons was observed formation of disulfide derivatives:
to depend on the ionic strength of the medium
according to the Brönsted equation (i.e., the Debye-
Hückel relation; eq 36). Extrapolation to zero ionic
strength gave a bimolecular rate constant of k0 = (2.5
± 0.5) × 109 M–1 s–1. The rate constant measured The oxidation of glutathione was fist order in
for reduction of metatungstate by CH3C•HOH, k = [CoIIIW125–] and [RSH] and inverse order in [H+]. A
(2.2 ± 0.4) × 108 M–1 s–1, was similar to that rate expression consistent with these observations
measured by Papaconstantingu (see Table 6). was derived for a mechanism involving acid dissocia-
In Rome, Baciocchi161 and co-workers later used tion of RSH and rate-limiting oxidation of RS– by
several pulse-radiolysis methods to measure the rates CoIIIW125–. The oxidations of L-cysteine and mercap-
of reduction of CoIIIW125– (E° = 1.01 V vs NHE) by a toacetic acid were first order in [CoIIIW125–], second
variety of acyclic and cyclic carbon-centered radicals order in [RSH], and inverse second order in [H+]. In
(alkyl, α-hydroxyalkyl, α-alkoxyalkyl, α-amidoxy- the oxidation of β-mercaptoethylamine, parallel paths,
alkyl, and benzyl), which, instructively, included first and second order in thiol, and an inverse
those previously studied by Papaconstantinou. dependence on [H+] were observed. Proposed mech-
Considering the high positive reduction potential anisms accounted for the [H+] dependencies by
of CoIIIW125–, rates of reduction of CoIIIW125– by including the acid dissociation of RSH and for the
hydroxyalkyl and carboxyl radicals were in line with dependencies on [RSH] by including reactions of RS–
those reported by Papaconstantinou (above).152–155–157 (L-cysteine and mercaptoacetic acid), or of RS– or
In particular, Baciocchi found that the rates of RSH (β-mercaptoethylamine), with association com-
reduction of CoIIIW125– to CoIIW126– by simple α-hy- plexes between CoIIIW126– and RS– (i.e., reaction of
droxyalkyl radicals were all near the diffision- RS– or RSH with [(COIIIW12O405–)(RS–)]6–). Evidence
controlled limit; •CH2OH < CH3C•HOH < (CH3)2C•OH for the inclusion of alkali-metal cations in these
and rate constants ranged from 2.4 to 4.0 × 109 M–1 association complexes was reported.
146 Chemical Reviews, 1998, Vol. 98, No. 1 Weinstock

The catalytic oxidation of organic sulfides (R2S, R The nature of the radical-cation intermediates
= Me, Et, and n-Bu) by vanadomolybdophosphates formed upon reaction of the phenyl sulfides with
(H3+n [PVn Mo 12–n O 40 ]( 3+ n )–, n = 1–6) and oxygen was CoIIIW126– was central to interpretation of reactivity
reported by Kozhevnikov, Matveev, and co-workers data.170 Thus, in addition to studying the reactivity
in 1979.164 Highly selective oxidations of R2S to of the sulfur-centered radicals, some of the factors
mixtures of R2SO and R2SO2 (but no other products) responsible for the formation of either carbon- or
were obtained with conversions of 100%. In stoichio sulfur-centered radicals were also clarified.169,170
metric reactions between several different POM In an initial reactivity study, the mechanism of
anions and sulfides in water, the following order of C-S bond cleavage in 1-phenylethyl phenyl sulfide
reactivity was found: Me2S p Et2S > n-Bu2S. The radical cation (eq 117) was investigated by determin-
monovanadium anion [PVMo11O40]4– was inactive in ing the stereochemistiy of the products of the reaction
water but quite effective in acetone, a result that may of the parent sulfide with CoIIIW125–. In question was
have been due to the degree of protonation, and hence whether the C-S bond cleavage occurred via unimo-
reduction potential of the anion in the two solvents, lecular (SN1) or bimolecular (SN2) mechanisms.
or to solvent-dependent differences in the kinetics of Initially, reaction of 1-phenylethyl phenyl sulfide
speciation or in the stabilities (formation constants) with CoIIIW125– in AcOH/H2O (70:30) at 50 °C gave
of potentially active speciation products. The authors the corresponding sulfur-centered radical:
argued that the absence of products of R-S bond
rupture implied that freely diffusing R2S•+ radicals
were not formed during the multielectron oxidation
process.
The utility of the CoIIIW126– anion as a probe in the
study of electron-transfer reactions is highlighted by
the use of this anion by Baciocchi to elucidate the
mechanisms of chemical and biochemical oxidations After completion of the reaction, products resulting
of a range of aryl sulfide compounds.165–170 In addi- from both C-H bond cleavage (Scheme 10, path a)
tion to its fundamental value, study of the reactivity and C-S bond cleavage (Scheme 10, path b) were
of the free-radical and radical ion intermediates found. As expected, the deprotonation pathway (path
generated by oxidation of these sulfur-containing a) became more important when base (KOAc) was
compounds is pertinent to processes ranging from the added.
industrial to the purely biological.165
The reactivity of simple alkyl sulfide radical cations Scheme 10
is complicated by the propensity of these ions to react
with their neutral parent compounds to form rela-
tively stable two-center-three-electron sulfur-sulfur
bonds.169 Radical cations derived by l-e- oxidation
of aromatic sulfides, however, have a much lower
tendency to “dimerize” via the formation of sulfur-
sulfur bonds. Thus, data pertaining to the reactivity
of aryl sulfide radical cations is more easily inter-
preted than that obtained from reactions of alkyl
sulfides. For this reason, and because of the lack of
previous attention paid to aryl sulfide oxidation
reactions, Baciocchi studied the oxidations of a
variety of benzyl phenyl sulfide compounds.
Depending on the nature of the substituents on the
benzylic phenyl ring (Ar in Scheme 9), however, l-e-
oxidation of these compounds gave either carbon-
centered cation radicals (Scheme 9, path a) or sulfur- More significantly, when the reaction was carried
centered radicals (Scheme 9, path b). out using enantiomerically pure (R)-(+)-1-phenyl-
ethyl phenyl sulfide, racemization was observed,
Scheme 9 along with modest (16%) inversion of configuration.
These results indicated that the C–S cleavage reac-
tion took place by two pathways. The simplest
possibility was that racemization occurred via a
unimolecular pathway and inversion via a competing
bimolecular route. However, several lines of evidence
suggested that unassisted cleavage of the C-S bond
led first to a contact radical-cation pair and then to
a solvent-separated radical-cation pair. Inversion
of configuration was attributed to nucleophilic attack
by solvent on the contact radical-cation pair; race-
mization resulted when the contact radical-cation
pair separated prior to reaction of the benzylic cation
with solvent (Scheme 11).
Electron-Transfer Reactions of POMs Chemical Reviews, 1998, Vol. 98, No. 1 147

Scheme 11 Olatunji and Ayoko 174 observed that the rates of


oxidations of thiourea, (H2N)2CS, and 1,1,3,3-tetram-
ethyl-2-thiourea, (Me2N)2CS (2.0–20 mM), were in-
dependent of [H+] at pH values well below the pK b
values of the substrates (Chart 5).
Chart 5. Structures of Thiourea and
1,1,3,3-Tetramethyl-2-Uourea

Reactions of phenyl sulfides with CoIIIW125– were


also used to help clarify a long-standing question
regarding the mechanism of oxidation of sulfides by The reactions were carried out at high ionic strength
enzymatic oxidation catalysts such as peroxidases (µ = 1.0 M, NaClO4) and at pH values sufficiently
(horseradish peroxidase, HRP)168 and microsomal low ([H+] = 0.2–0.8 M) that the substrates were
cytochrome P 450167,170 and by biomimetic catalysts assumed to be monoprotonated. On the basis of
such as meso-(tetraphenylporphyrinato)iron(III) chlo- reaction stoichiometries and UV-vis spectroscopic
ride (TPPFeIIICl).167,170 It had been generally ac- analysis of independently synthesized disulfide stan-
cepted that oxidation of sulfides (R2S) to sulfoxides dards, the reaction with CoIIIW125– was represented
(R2S=O) by hemoproteins was initiated via transfer as
of an electron from the sulfide to the Fe(IV) oxo
porphyrin radical cation (Por•+-FeIV=O) (path a,
Scheme 12). According to this mechanism, the
resultant sulfur-centered radical cation, R2S•+, was
oxidized to the sulfoxide by reaction with the neutral
Fe(IV) oxo porphyrin (Por-FeIV=O; oxygen-rebound At 25 °C under pseudo-first-order conditions
mechanism) (path b, Scheme 12). Alternatively, ([CoIIIW125–] = 0.2 mM), the oxidation of monoproto-
however, the sulfide might be oxidized to the sulfox- nated thiourea was first order in oxidant and sub-
ide by direct oxo transfer from Por•+-FeIV=O (path c, strate (kobs = 12.50 ± 0.03 M–1 s–1), while monopro-
Scheme 12). tonated (Me2N)2CS was oxidized by two parallel
pathways, first and second order in substrate.
Scheme 12 One mechanism proposed for the oxidation of
monoprotonated thiourea included the formation of
an association complex, followed by electron transfer
and rapid radical coupling:

Where reaction of sulfide with CoIIIW125– gave


sulfoxide, an additional mechanism, hydrolysis of the
first-formed radical cation, R2S•+, to give R2S•(OH)
followed by l-e- oxidation by a second equivalent of
COIIIW125–, was proposed.170
By comparing the reactivities of phenyl sulfides
with CIIIW125– and HRP, Baciocchi confirmed that Assuming 1 p Ki p[((R2N)2CS)(H+)], these elementary
reactions with HRP occurred via an oxygen-rebound steps gave the rate expression
mechanism and measured the rate of the oxygen-
rebound step.168 In addition, the products obtained
by reactions of phenyl sulfides with CIIIW125–, with
enzymes, and with biomimetic oxidants were inter-
preted as evidence that reactions with microsomal
cytochrome P 450 and (TPPFeIIICl) most likely oc-
curred via oxo transfer.167,170 However, recent kinetic where [CoIIIW125–]T = [CoIIIW125–] + [CoIIIW125–)-
studies have shown that apparent rate constants for (R2N)2CS)(H+)]4– and Ki pket = 12.50 ± 0.03 M–1 s–1.
oxygen rebound by cytochrome P 450 can be incred- In an effort to determine whether the reaction
ibly fast and that nonsynchronous concerted mech- occurred via an outer-sphere electron-transfer mech-
anisms are possible. 171–173 Given this apparent con- anism, evidence for the existence of a linear free-
tinuum between oxygen-rebound and oxo-transfer energy relationship between log kobs and E° for this
mechanisms, conclusions arrived at through product and a set of similar reactions of CoIIIW125– with
analysis studies, however elegantly conceived, must organic and inorganic substrates (glutathione,162,163
be viewed with a degree of circumspection. N2H5+,175 and HNO2176) was sought. Although a
148 Chemical Reviews, 1998, Vol. 98, No. 1 Weinstock

reasonably good correlation was found (linear regres- reaction of the dehydrated intermediate, H5[PV2-
sion coefficient = 0.98), the authors claimed that the Mo10O40], with THT. Subsequent reaction of THT
slope of the line (–8.6 V–1) was inconsistent with with water and a second equivalent of POM was
either classical adiabatic outer-sphere electron trans- assumed to occur rapidly. Assuming that the con-
fer within donor-acceptor “precursor” complexes centration of the dehydrated intermediate, H5[PV2-
(expected slope = +8.5 V–1; i.e., a positive value)3 or Mo10O40], remained constant, a general rate expres-
reversible, endergonic outer-sphere electron transfer sion was derived using a steady-state approximation.
within ion-paired association complexes (slope = At high [H2O] the general rate expression gave the
–16.9 V–1).54 Confronted with this discrepancy, the observed rate law, rate = kobs[THT] [POMtot] [H2O]–1.
authors concluded that although the reaction did not The steady-state rate expression was also consistent
comply with Marcus’s theory, the substitutionally with the fractional order of reaction with respect to
inert nature of [CoIIIW12O40]5– argued in favor of an [THT] observed at low [H2O].
outer-sphere mechanism. It is well-established that water present within the
It appears, however, that the experimental data crystalline lattices of solid POM acids and salts plays
are entirely in line with theoretical expectations; an important role in the heterogeneous-phase cataly-
based on accepted convention, the sign of the slope sis of reactions of organic substrates in nonaqueous
of the plot of log kobs vs E°, where E° values are re- media.144 It is also well-known that the reactivity of
duction potentials of electron donors (i.e., the reduc- organic-soluble d-electron-containing transition-metal-
tion potentials of the 1e–-oxidized forms of the substituted POM anions in nonaqueous media is
donors) should be negative (see section II.D). Because strongly influenced by the presence of water or of
Ayoko and Olatunji, the authors of the data used in other oxygen- or nitrogen-donor ligands.178,179 In the
this analysis, appear to be first-rate experimentalists, latter cases, the donor ligands compete with sub-
their data should be regarded as highly reliable. It strates for occupation of the sixth coordination site
is gratifying that an elegant conclusion might be on the d-electron-containing transition-metal cation.
drawn from it. However, as the present report demonstrates, small
In 1994, Hill studied POM (H5[PV2Mo10O40]•(H2O)n) concentrations of water can also influence the non-
catalyzed oxidations of thioethers, RSR' (including aqueous chemistry of organic-soluble POMs that do
the mustard gas analogue 2-chloroethyl ethyl sulfide, not appear to possess open addendum-atom coordi-
CH2(Cl)CH2SCH2CH3 by tert-butyl hydroperoxide nation sites. As such, the results of Hill’s study could
(TBHP) in acetonitrile.177 The larger context for the have broad technological implications for the use of
detailed mechanistic studies reported was the use of organic-soluble POM catalysts in nonaqueous sol-
POMs as catalysts for rapid destruction of toxic vents.
agents.
Reactions of thioethers with TBHP in the presence
of H5[PV2Mo10O40]•(H2O)n at room temperature gave V. Oxidation of lnorganic Electron Donors by
sulfoxides, RS(O)R', in remarkably high yields POM Anions
(>99.9%) with no evidence of overoxidation to sul-
fones, RS(O2)R'. The high selectivity was traced to A. Introduction
the mode of catalysis: The thioethers were directly
oxidized by H5[PV2Mo10O40. The role of TBHP in the Experimental rate laws have been determined for
catalytic cycle was simply to reoxidize the reduced reactions of CoIIIW125– and other POMs with a wide
POM anions, H6[PV2Mo10O40]. Kinetic and mecha- variety of inorganic electron donors. These studies
nistic studies thus focused on the stoichiometric are summarized in tabular form in Table 7.
oxidation of a model thioether compound, tetrahy- In addition to the reactions summarized in Table
drothiophene (THT), by solutions of H5[PV2Mo10O40]• 7, numerous examples of homogeneous oxidations
(H2O), in acetonitrile (the nature of acetonitrile catalyzed by multicomponent systems consisting of
solutions of H5[PV2Mo10O40]•(H2O)n was discussed in Pd(II) salts, vanadomolybdophosphate complexes
section IV.B). (H3+n [PVn Mo 12–n O 40 ] (3+n )–), substrate, and oxygen
In the course of these studies, an inverse depen- have been reported.86,87,113,114,145,180–190 Generally, the
dence of reaction rate on the concentration of H20 POM anions play a role analogous to that of Cu salts
was observed. When water was present at concen- in the Wacker process. During the catalytic cycle,
trations of 2.2 × 10–1 M (~0.4 vol %) or greater, the Pd(II) is reduced by the substrate (SubH2, eq 123),
rate law, rate = kobs[THT] [POMtot] [H2O]–1 ([POMtot] and the POM anion catalyzes reoxidation of the
is the total concentration of catalyst, H5[PV2Mo10O40]• reduced-Pd species. After electron transfer from the
(H2O) n, where n is zero or greater) was observed. reduced-Pd species to the POM anion (eq 124), the
However, as [H2O] was decreased from 0.4% to the resultant heteropoly blue is oxidized by O2 (eq 125).
experimental limit of 1.5 × 10–2 M (a limit imposed
by hydration of H5[PV2Mo10O40] and the hygroscopic
nature of the solvent), the order of reaction with
respect to [THT] decreased to near 0.5.
Inhibition by water (the inverse dependence on
H2O]) was attributed to unimolecular loss of water
from H5[PV2Mo10O40]•(H2O)n In the proposed mech-
anism, loss of water occurred prior to bimolecular
Table 7. Experimental Rate Laws for Oxidations of Inorganic Electron Donors by [CoIIIW12O40]5–(CoIIIW) and by Other POM Anions
150 Chemical Reviews, 1998, Vol. 98, No. 1 Weinstock

These reactions sum to the net oxidation of the reactivity of the molybdate complexes followed the
substrate by O2: order (with respect to the number of vanadium(V)
ions, n) n = 0 p n =1 > n = 2 > n = 3 > n = 6.
Notably, the order of the reduction potentials of the
anions at pH 3 (M = W > M = Mo and n = 1 > n =
The global kinetics of these multicomponent sys- 2 > n = 0) did not correlate with their reactivities.
tems and the effects of specific variables on overall Pseudo-unimolecular kinetics plots (ln of absorbance
reaction rates have been studied in detail.86 How- vs time) were obtained for the molybdates with n =
ever, little if any detailed information concerning 0, 1, and 2. However, the dependence of rate on
oxidation of the reduced-Pd species by POM anions [POM] demonstrated saturation kinetics or some
is available. One reason is that the reduced-Pd other equilibrium-based phenomenon. Namely, solu-
species formed during catalyst-turnover conditions tions of Na2MoO4, adjusted to pH 3 (i.e., solutions of
(usually86written as Pd(0)) are unstable and poorly isopolymolybdates), exhibited kinetic behavior identi-
defined. cal with that of [PMo12O40]3– solutions. The authors
The potentially commercially useful POM-cata- suggested that the kinetically active oxidants were
lyzed oxidation of hydrogen sulfide to elemental actually isopolymolybdates, products of speciation
sulfur was first explored using vanadomolybdophos- (i.e., reaction with water and electrolyte) of the
phate complexes, H3+n [PVn Mo 12–n O 40 ] (3+ n )–, where n vanadium-free and vanadium-substituted heteropoly-
= 1–6. In later studies, oxidations of hydrogen molybdate precursors.
sulfide by well-defined heteropolytungstate anions In 1989 Kuznetsova, Chernyshova, and Maksimo-
were investigated. Available information concerning vskaya210 studied the oxidation of H2S at 20 °C
the mechanisms of oxidation of hydrogen sulfide by by the heteropolytungstates earlier examined by
heteropolymolybdate and tungstate anions is sum- Kozhevnikov.209 One intellectually compelling goal
marized below. In addition, due to its theoretical of the study was to determine the effects of posi-
importance and because of its value in illustrating tional isomerism on rate. Reactions of the anions
intellectually substantive issues pertinent to the role [PVn W 1 2 –n O 40 ] (3+n )–, where n = 1, 2 (neighboring V
of donor-acceptor ion pairing in outer-sphere electron- atoms and mixture of positional isomers), 3 (α-1,2,3
transfer reactions (see section II.D), the luminescence isomer and mixture of isomers), and 4 (mixture of
quenching of Ru(II) complexes by well-defined and positional isomers), were carried out over a range of
kinetically inert heteropolytungstates is presented in pH values (pH 1.5–3 for n = 1 or 2 and 2–5 for n =
detail. 3 or 4) under pseudo-first-order conditions of excess
POM. Critically, both 31P and 51V NMR were used
B. Inorganic Electron Donors to confirm that the heteropolytungstates (and their
isomeric forms) remained intact during the reac-
Balanced equations and experimentally deter- tion.
mined rate laws for oxidations of inorganic electron Although most of the kinetic data was obtained at
donors by CoIIIW125–, and by other POMs, are listed high ionic strengths (µ = 0.25–0.5 M, Na2SO4), no
in Table 7. With few exceptions, the studies were dependence of rate on ionic strength was observed
carried out under conditions of high ionic strength. over a dramatic range of electrolyte concentrations
In general, the phenomena noted in the oxidation of (0–2.0 M Na2SO4). At pH 2.6 the reaction was zeroth
organic substrates (section IV, above), such as acid- order in the n = 1 or 2 anions and, provided the
base equilibria, ion pairing, the formation of donor- [POMI was not excessively large, first order in the n
acceptor complexes, and catalysis by alkali-metal = 3 and 4 anions. At higher pH values (4.5–5.0) and
cations, were observed in the studies reported here sufficiently high POM concentrations, the reactions
as well. were zeroth order in the n = 3 and 4 anions, as well.
At low pH values (1.3–3.0), an inverse dependence
C. Hydrogen Sulfide of rate on [H+] was observed for the n = 1 and 2
anions. More complex dependencies on [H+] were
In 1975 Matveev disclosed208 the catalytic oxida- observed at higher pH values.
tion of hydrogen sulfide (H2S) to elemental sulfur On the basis of analysis of the kinetic data, the
(S8) by vanadomolybdophosphate anions, H3+n [PVn - authors proposed that the rate-limiting step in the
Mo12–n O40](3+ n )–: reactions involved the unimolecular decomposition of
hydrogen-bonded structures of the form shown in
Chart 6.
The kinetics of the stoichiometric oxidations of Chart 6. Hydrogen-Bonded Intermediate in the
aqueous solutions of H2S by [PVn M 1 2 –n O 40 ] (3+ n )– (M Oxidation of HS–
= Mo, n = 0–6; M = W, n = 0, 2, and 3) were later
investigated by Kozhevnikov and Kholdeeva.209 The
pK a of the first acid dissociation of H2S is 7.0, and
the kinetic studies were carried out at pH 3 ([H2S]/
[HS–] = ~104) in Na2SO4/NaHSO4 buffer (µ = 2.0 M)
using a stopped-flow apparatus.
At 25 °C the heteropolytungstates were inactive on
the stopped-flow-experiment time scale, while the
Electron-Transfer Reactions of POMs Chemical Reviews, 1998, Vol. 98, No. 1 151

Given that the anions studied were stable over


different pH ranges, a suitable rate expression for
the calculation of constants for a potentially common
rate-limiting elementary-reaction step might have
required a rate expression that included (1) the pK a The value of kobs found using the initial rate method
of H2S, (2) the formation of donor-acceptor ion pairs, was kobs = 79.3 ± 7.1 M–2 s–1. However, nonlinear
and (3) the pK a of H2S within the donor-acceptor ion least-squares fit of absorbance vs time data to the
pairs. However, the entire pH range over which all appropriate integrated third-order kinetic expression
the reaction were studied (~1.5–5) was sufficiently gave kobs = 96 ± 43 M–2 s–1, and as the reaction
below the pKa of H2S (pKa = 7.0), such that acid progressed, systematic deviations in kobs increased.
dissociation of H2S was not included in the proposed The large experimental error and systematic devia-
reaction sequence: tion with reaction time were attributed to the com-
plex chemistry of the aqueous sulfide solutions and
to disparate reactivities of partially oxidized sulfur
intermediates.
The [H+] dependence of the initial rate, investi-
gated over a pH range of 7.5 to 9.5 at constant ionic
strength, was complex. Initial rates rose slightly
from pH 7.5 to ~8 and dropped by a factor of 5 from
pH 8 to 9.5. Deprotonation of the POM anion was
From a rate expression based on eqs 128–130, values one explanation tentatively proposed to account for
of kslow were calculated for the anions (n = 1–4) the [H+] dependence. Although repeated acid-base
studied. One significant result was clear: For n = titrations of solutions of K14[NaP5W30O110] failed to
3, kslow for oxidation by the α-1,2,3 isomer was greater exhibit a clear endpoint, the POM anion appeared
(1.3 min–1 at pH 4.5) than kslow for oxidation by the to undergo deprotonation somewhere in the 7.4–9.5
mixture of n = 3 isomers (0.6 min–1 at pH 5.3). This pH range.
result was rationalized in terms of the greater On the basis of the experimental rate law and the
basicity of the µ-2 oxygen atoms coordinated to [H+] dependence of initial rates, the following el-
neighboring V(V) ions in the α-1,2,3 isomer (Chart ementary reaction steps were proposed:
6, M = V(V)). It should be noted that very few such
investigations of the effects of polarization (i.e., loci
of higher electron density) within soluble POM anions
have been reported.
In separate work aimed at the development of
catalytic processes for the oxidation of hydrogen
sulfide by oxygen, Kuznetsova211 reported that the
end products of the reaction of [PFeIII(OH2)W11O39]4–
with H2S at pH 4–5 were [PFeII(OH2)W11O39]5– and
elemental sulfur.
Harrup and Hill later reported the catalytic oxida-
tion of H2S to S8 by a variety of transition-metal-
substituted and mixed-addenda heteropolytung-
states and oxygen. 212,213 In conjunction they studied
the stoichiometric oxidation of HS– in aqueous
solutions of the potassium salt of the remarkably
alkali-stable214 Pope–Jeannin–Preyssler anion,
K14[NaP5W30O110],215 at pH 8.5, (0.025 M Borax buffer,
µ = 0.310 M)213 (eq 131).

Ion pairing between [NaP5W30O110]14– and the alkali-


The rates of oxidation of the sulfur-redox interme- metal cations present, although not addressed ex-
diates (polysulfide (Sn 2–), S2O62–, S2O32–, and S8) were perimentally, was mentioned as a likely possibil-
also examined. Relative rates of reaction were HS– ity.
(0.96), Sn 2– (22.2), S2O62– (159), S2O32– (0), and S8 (0). Harrup and Hill212 observed a different rate law
To avoid the oxidation of sulfur-redox intermediates, in the oxidation of HS– by the tetramethylammonium
some of which were oxidized more rapidly than HS– (TMA) salt of the Nb3O13-capped Wells-Dawson
itself, conversions of HS– were kept to low values. anion [Nb3P2W15O62]9–.216 As with [NaP5W30O110]14–,
Reactant concentrations were also kept low (1.0–3.0 the reaction was carried out in Borax-buffered aque-
and 0.7–10.0 mM, respectively, for [NaP5W30O110]14– ous solutions with low reactant concentrations and
and HS–). Based on log-log plots of initial rates, the low conversion rates. As before, the stoichiometry
reaction was first order in [HS–] and second order in shown in eq 131 was verified. Unlike in the previous
[(NaP5W30O110)14–]: study, however, log-log plots of initial rates revealed
152 Chemical Reviews, 1998, Vol. 98, No. 1 Weinstock

that the reaction was first order in both [HS–] and


[Nb3P2W15O62)9–]:

The possibility of 2-e– reduction of the [Nb3W15O62]9–


anions was discounted on the basis of thermodynamic
considerations. In addition, log-log plots of the Figure 6. Static and dynamic energy-transfer quenching
dependence of rate on [H+] (pH range of ~7–9) were of
3
the luminescence emission of [Ru(bpy)3]2+*. D, 1D*, and
linear, but a slope of 0.4 indicated that the role of D* represent the ground, lowest-excited singlet, and
protons in the reaction was complex. Acknowledging lowest-excited triplet states of [Ru(bpy)3]2+, and τ° is the
oversimplification, the following rate-limiting electron- lifetime of the luminescence emission in the absence of the
quencher, A. Kip is an ion-pair association constant.
transfer step was proposed:
emission at the same rate at which it shortens the
emission lifetime:

D. Luminescence Quenching of Electronically


Excited [Ru(bpy) 3 ] 2+ Cations As a result, plots of το/τ or I°/I vs [A] are linear and
coincident; either can be used to determine k4 .
Luminescence from photochemically generated elec- Static quenching, if present, does not alter the ratio
tronically excited states of [Ru(bpy)3]2+ cations (bpy of k4 /k3. Hence, the lifetime of the luminescence
= 2,2'-bipyridine) can be quenched by a number of emission remains unchanged; eq 140 still applies.
mechanisms and pathways.217–222 One luminescence- Static quenching does, however, attenuate the steady-
quenching process, energy transfer to quencher A, stats intensity of the luminescence emission. In this
where D* represents the photoexcited Ru(II) complex case (provided that the dynamics of photoexcitation
[Ru(bpy)3]2+• (eq 139), is instructive of the principles of D in D:A parallel that of D and k5 is fast), the
involved. emission intensity decreases at a rate determined by
both k4 and Ki p :

The luminescence emission of [Ru(bpy)3]2+* can be


quenched by both static and dynamic energy-transfer
mechanisms (Figure 6). 218 Static quenching is as- When both static and dynamic quenching are present,
sociated with the quenching of luminescence inten- the plot of I°/I vs [A] (Stem–Volmer plot of emission
sity within (D:A)* ion pairs (Kip and k5 in Figure 6), intensity) is curved and deviates from the line
while dynamic quenching of both emission lifetime described by eq 140. Mathematical comparison of the
and intensity occurs during the collision encounter plot of I°/I vs [A] (eq 142) to the plot of το/τ vs [A] (eq
between D* and A (k1 or k4). Static quenching of 140) provides Ki p .
luminescence intensity is observed at the wavelength In the luminescence quenching of photoexcited
of the light emitted during stationary irradiation; [Ru(bpy)3]2+ cations by [PMnII(OH)W11O40]6– and
lifetime measurements require rapid-time-scale moni- [SiCoII(OH2)W11O40]6–,60 a much more complex situ-
toring of the decay in luminescence emission after ation pertained. 61,223 Namely, the [Ru(bpy)3]2+ and
photoexcitation. POM anions were highly ion paired (K ip values
In the absence of static or dynamic quenching, the ranged from ~103 to nearly 105 M–1, depending on
lifetime (τ°) of the luminescence emission induced by ionic strength of the solution).61 Both static and
photoexcitation of [Ru(bpy)3]2+ is determined by k3 . dynamic oxidative electron-transfer quenching were
In the presence of A, dynamic quenching of either observed and in both cases, occurred within ion pairs
the initially generated singlet-excited-state complex, (Figure 7).
1[Ru(bpy)3]2+* (k 1), or3 of the lower energy triplet- One of the challenges encountered in studying
excited-state complex, [Ru(bpy)3]2+* (k 4), can occur. electron transfer between reactants that form stable
218
However, under certain conditions, k1[A] ^ k2, so donor-acceptor ion pairs is to accurately determine
that dynamic quenching is described by k4. Under rate constants for the “intramolecular” electron-
these conditions, the lifetime of the emission is transfer step (e.g., kfet in Figure 7). In these cases,
shortened to an extent dictated by the product values for kfet must usually be extracted from ko b s
k4[A]. Plot of the ratio τ°/τ vs [A], τ being the values, which are often composites of a number of
emission lifetime in the presence of A, gives a straight rate and equilibrium constants (see discussion of
line (Stern–Volmer plot of emission lifetime) whose Scheme 2 in section II.D). In addition, if the reaction
slope is proportional to k4 : is diffusion controlled (i.e., kfet (in units of s–1) > kd -
[Donor][Acceptor]; Scheme 2 and Figure 7), ko b s
values might provide only a lower limit on kf e t .
Ballardini, Gandolfi, and Balzani, however, devised
In the absence of static quenching, dynamic quench- a strategy for direct determination of the rate con-
ing decreases the intensity of the luminescence stant for intramolecular electron transfer between
Electron-Transfer Reactions of POMs Chemical Reviews, 1998, Vol. 98, No. 1 153

lay the key to what followed. Able to determine K,


and recognizing that kq ≈ kd, Ballardini and co-
workers sought to determine whether the ionic-
strength dependence of kq obeyed the Debye–
Smoluchowski (DS) equations37 (see discussion of the
derivation of eq 1 in section IIA). The DS equation
provides kd as a function of ionic strength for spheri-
cal reactants of specific size and charge freely dif-
fusing in a dielectric medium. Ballardini et al. found
that for ionic radii of 7.0 and 5.5 Å for the Ru(II)
Figure 7. Static and dynamic luminescence quenching of complex and POM, respectively, kq did in fact obey
photoexcited [Ru(bpy)3]2+ (A2+) cations by POM anions, Q 6–
(Q6– = [PMnII(OH)WllO40]– or [SiCoII(OH2)W11O40]6–), via
the DS equation. Because the DS equation was
oxidative electron transfer within *A2+:Q6– ion pairs. The obeyed, it followed that the Eigen-Fuoss (EF) equa-
reduced POM anions, Q7–, are heteropoly blues, formed by tion, used to calculate KEF ( = kd /k –d) and based on
1-e– reduction of the tungsten oxide frameworks of Q6–. the same theoretical model as the DS equation,
Static and dynamic pathways are indicated by dotted should provide an accurate estimate of the Kexp values
lines. obtained from Stern–Volmer plots. This, however,
was not the case.
Ru(II) and W(VI) in [(RuII(bpy)3)*(SiCoII(OH2)- Throughout the ionic-strength range over which
W11O40)4– ion pairs (kfet in Figure 7).60 the DS equation was obeyed by kq, values of KEF were
As part of a larger study,6l Ballardini and co- smaller than Kexp by a constant factor. This result
workers discovered that in solutions of [Ru(bpy)3]2+ suggested that the effective encounter distance might
and [SiCoII(OH2)W11O40]6–, at moderate ionic strength be smaller than the sum of the radii of the reactants.
(µ = 0.01 M, LiCl), rates of static quenching were However, decreasing the encounter distance in the
much larger than those of dynamic quenching (see Eigen-Fuoss calculation did not bring KEF values in
caption to Figure 7). Rather than a single composite line with experimental observation. A good fit to
curve, the exponential decay in luminescence inten- experimental data was obtained, however, by mul-
sity observed after pulse excitation was bimodal, each tiplying KEF (calculated using an encounter distance
mode corresponding to one of the two quenching of 12.5 Å) by a constant.
pathways. Thus, the decay constant associated with To understand how the authors rationalized this
static quenching could be individually discerned. In result, it is essential to note that Kexp values, derived
addition, thermodynamic arguments dictated that from analysis of Stern–Volmer plots, represented the
staticenergy-transfer quenching (1/τ°ip, Figure 7) was equilibrium constants for formation of the ion pairs
ineffective relative to the static-electron-transfer involved in static quenching. The kq values, on the
pathway, kfet. As a result, the kobs values associated other hand, were derived from the emission-lifetime
with static quenching depended only upon kfet, kd, and measurements associated with dynamic quenching.
k–d. Hence, the authors proposed that the kq (≈ k d) values,
By using the decay in luminescence intensity which obeyed the DS equation, represented the rate
attributed solely to static quenching as the experi- of formation of “regular” encounter complexes with
mental observable, solution of the integrated rate encounter distances of ~12.5 Å The Kexp values,
expression for static quenching gave kfet = 8.5 × 107 however, associated with intensity quenching within
s–1, kd = (3.5 ± 0.1) × 1010 M–1 s–1, k–d = (1.6 ± 0.5) preexisting ion pairs (Figure 7), were larger than
× 106 s–1, and K = kd /k–d = ~2 × 104 M–1. Notably, those predicted by the Eigen-Fuoss equation because
kfet values were independent of the concentration of of penetration of the POM anion into a pocket of the
the quencher, [SiCoII(OH2)W11O40]6–. cation structure.
More generally, information regarding the detailed On the basis of ab initio calculations, similar
structure of the [(RuII(bpy)3)*(SiCoII(OH2)W11O40)]4– interpenetration of coordination spheres within
and [(RuII(bpy)3)*(PMnII(OH)W11O40)]4– ion pairs was intimate ion pairs was once proposed for the
inferred from theoretical evaluation of the observed FeIII/II(H2O)6]3+/2+ electron self-exchange reac-
effects of ionic strength on the constants (K = kd /k – d , tion. 53
Figure 7) associated with their formation.223 The
observed rate constant for dynamic quenching, kq
(Figure 7), is equal to the product, Kkfet, where K = VI. Reduction of Halogenated Alkanes by
kd /k–d. Values of kq were obtained from Stern– Reduced POM Anions
Volmer plots of το/τ vs [A], while association con-
stants, K, were derived from combined lifetime and A. Carbon Tetrabromide
intensity-quenching experiments. Both kq and K
were determined at several ionic strengths (µ = 4 × Carbon tetrabromide (CBr4) was reduced by the
10–3 to 0.5 M), using the Co(II)- and Mn(II)-substi- one-electron heteropoly blue K7[CoIIW12O40] (~0.2
tuted heteropolytungstates and Ru(II) complexes mM) at pH 7 (buffered by tris(hydroxymethyl)-
containing several ligand types (bpy, 4,4'-Cl2bpy, aminomethane, TRIS) in mixtures of MeCN and H2O
etc.). containing low concentrations of added electrolyte
At constant ionic strengths, kq values were inde- (1–200 mM Li+, Na+, K+, Cs+, Me4N+, and Bu4N+).224
pendent of ∆G°, a strong indication the process was The only products found were CHBr3 (77%) and C2-
diffusion controlled (kq ≈ kd; see section II.D). Herein Br6 (1%). Yields were based on consumption of
154 Chemical Reviews, 1998, Vol. 98, No. 1 Weinstock

[CoIIW12O40]7–, and control experiments confirmed relative abilities of specific cations to participate in
that the dimer C2Br6 did not result from uncontrolled the formation of donor–M+–acceptor association
photochemical processes. The stoichiometric reaction complexes.52
is shown in eq 143. On the basis of product analyses, kinetic data, and
pertinent control experiments, a mechanism for the
reduction of CBr4 by [CoIIW12O40]7– was proposed:

The potentials of the 1- and 2-e– reductions of


[CoIIW12O40]6– as a function of pH were mea-
sured by cyclic voltammetry. Due to protonation
of [CoIIW12O40]7– (or of both [CoIIW12O40]6– and
[CoIIW12O40]7– in the mixed-solvent system), addition
of H+ increased the electrochemical potential of the
[CoIIW12O40]6–/7– couple. (In water, no protonation of
[CoIIW12O40]6– occurs even at pH 0.)63 At low pH
values, positive shifts in the potential of the
[CoIIW12O40]6–/7– couple led to decreases in rates of
reduction of CBr4.
Similarly, comparison of the relative rates of reac-
tions run using 10 mM concentrations of added
electrolyte suggested that ion pairing slowed the rate
of reduction of CBr4 (numbers in parentheses are Implied in this reaction sequence was the assump-
relative rates): Me4N+ (3.0) > Bu4N+ (2.2) > Li+ (1.0) tion that in contrast to aliphatic monohalides, the
≈ Na+ (1.0) > K+ (0.7) > Cs+ (0.5). Moreover, polyhalogenated compounds could form relatively
saturation kinetics was observed in the deceleration long-lived anion radicals (experimental data in sup-
of reaction rates upon incremental addition of port of this assumption is discussed in section VI.B,
Na+ClO4– (1.0–200 mM) to 0.2 mM solutions of K7- immediately below). Spin-trapping studies (ESR
[CoIIW12O40] (1.4 mM in K+). This behavior was spectroscopy of reactions run in the presence of the
consistent with formation of ion-paired species, spin trap N-tert-butyl- α-phenylnitrone, PBN) pro-
M +x M' + y [CoIIW12O40](7 – x – y ) – , where M = K, M' = Na, vided evidence for Br3C• radicals.225 Ion pairing,
and x + y was probably = 2,23 which shifted the although discussed, was not explicitly represented in
potential of the [CoIIW12O40]6–/7– couple to more the elementary steps of the proposed reaction mech-
positive values. anism.
Saturation kinetics was also observed in the ac- An expression, identical in form to the experimen-
celeration of reaction rates upon incremental addition tal rate law (eq 148), was derived by setting d[Br3C• /
of Me4N+ClO4– (0–100 mM) to 0.2 mM solutions of dt = 0 (steady-state approximation) and by assuming
K7[CoIIW12O40]. The electrochemical potential of the that k2 p k1 = kobs = 1.1 ± 0.3 M–1 s–1 p k –1 and
[CoIIW12O40]6–/7– couple (0.5–2.0 mM K6[CoIIW12O40]) that k3 was fast:
was measured by cyclic voltammetry, E° = (Eanodic –
E cathodic)/2, in 100 mM solutions of NaClO4 or Me4-
NClO4 buffered at pH 7 in 64% MeCN. E° values
(vs NHE) of –439 mV (NaClO4) and –485 mV (Me4-
NClO4) were observed. These values were consistent
with the acceleration in rate observed upon addition The rate of reduction of CBr4 by the two-electron
of Me4N cations: Shift in the potential of the heteropoly blue [CoIIW12O40]8– was also studied. The
[CoIIW12O40]6–/7– couple to more negative values by observed rate constant, k1 ( [ C o W ] 8 –), was compared to
addition of Me4N+ resulted in an increase in ∆G° for k1 (i.e., k1 ( [ C o W ] 7 – )) from eq 144 on the basis of two
reduction of CBr4. assumptions: (1) electron transfer from [CoIIW12O40]8–
In a related study, Carloni and Eberson measured to CBr4 was rate limiting, and (2) reduction of Br3C•
the effects of added cations on the chemical potential by [CoIIIIW12O40]8– was fast relative to its reduction
of the [CoIIIW12O40]5–/[CoIIW12O40]6– couple.52 The by [Co W12O40]7–. The ratio of experimental rate
addition of Me4N– shifted the potential of the constants, k1([CoW]8 – )/k 1([CoW] 7 – ), was equal to 4, while the
[CoIII/IIW12O40]5–/6– couple to more negative values, ratio predicted using Marcus theory was 3. The
while addition of Na+ shifted the potential in the relatively good agreement suggested that the rates
positive direction. In electron transfer between of both reduction reactions were controlled by activa-
[CoIIW12O40]7– and neutral CBr4, the kinetic impact tion energies associated with parallel outer-sphere
of added cations is due primarily to the effect of ion electron-transfer processes.
pairing between M+ and [CoIIW12O40]7– on ∆G° for
the electron-transfer step. However, Carloni and B. Halomethanes
Eberson noted that in other reactions, notably elec-
tron transfer between two charged species, the Eberson and Ekström225 also measured the rates
thermodynamic (and hence kinetic) effects of ion of reduction of a series of polyhalogenated alkanes.
pairing on ∆G° could be counterbalanced by the The following order of reactivity toward [CoIIW12O40]7–
Electron-Transfer Reactions of POMs Chemical Reviews, 1998, Vol. 98, No. 1 155

Table 8. Product Distributions and Yields for the Reductive Dehalogenation of Halocarbons by Reduced
Polytungstates a

was found: CBr4 > (CBrCl2)2 > CBr2Cl2 > C2Cl6 > ([PW12O40]4–) and 1.6 ± 0.05 electrons (3:2 [W10O32]6–/
CBrCl3 > CF3HClBr > CHBr3 > CCl4. Spin-trap [W10O32]5– mixture). Potentiometric titration with
studies (PBN) provided evidence for the formation of hydroxide (introduced as n-Bu4NOH) showed that the
intermediate radicals during reduction of the most- reduced species, [HPW12O40]3– and [H2W10O32]4–,
and least-reactive substrates, CBr4 and CCl4. Reduc- were protonated. However, the endpoint for titration
tion of CBrCl3 by [CoIIW12O40]7– (eq 149) obeyed the of the reduced decatungstate solution was less dis-
same rate law as that found for the reduction of CBr4 tinct due to the presence of the 1-e– heteropoly blue
(eq 148, above). [W10O32]5–. In kinetic studies of the haloalkane
reduction reactions, the progress of the reaction was
followed spectrophotometrically. Quantification of
the oxidation of [HPW12O40]3– was straightforward;
oxidation of mixtures of [H2W10O32]4– and [HW10O32]4–
to fully oxidized [W10O32]4– was accomplished by
Comparison of the rate constants for the two simultaneous observation of the spectral character-
reactions gave k1(CBr4)/ k 1(CBrCl3) = 80. The authors istics of all three species. Product distributions and
argued that if the mechanism of the electron-transfer yields are shown in Table 8.
reactions was dissociative (i.e., dissociation of a halide The products were dominated by those typically
anion upon reduction and no radical-anion interme- derived from disproportionation, coupling, or l-e-
diates in eqs 144 and 145), both CBr4 and CBrC4 reduction of free-organic-radical intermediates. Clear
would have reacted at the same rate. On the basis evidence for disproportionation and coupling was
of published thermodynamic data (enthalpies and seen in the 1:1 ratio of cyclohexane to cyclohexene
entropies of dissociation), the difference between the and in the formation of dimeric (C6H11)2 upon reduc-
reduction potentials of the two haloalkanes was tion of cyclohexyl bromide (Table 8, entry 4 and eq
calculated to be small: E° (CBr4) – E° (CBrCl3) = 150). The distribution of products obtained upon
+0.01 V. In addition, good linear relationships were
known to exist between the energies of dissociation
and (concomitant) reorganization, λ, of simple ali-
phatic halides. On the basis of Marcus analysis, it
was reasoned that if the electron-transfer step was
dissociative, rates of reduction of the two substrates
would be very similar. Hence, the relatively large
value observed for the k1(CBr4)/ k 1(CBrCl3) ratio was
seen as evidence for relatively long-lived radical-
anion intermediates. reduction of CBrCl3 (Table 8, entry 2) was almost
identical with that observed by Eberson and Ek-
C. Carbon Tetrachloride, Halomethanes, and ström225 after its reduction by the 1-e– heteropoly
Haloalkanes blue [CoIIW12O40]7–: CHCl3 (72%), C2Cl6 (2%), and
CHBrCl2 (< 1%).
The reductive dehalogenation of the haloalkanes Initial rates of reduction of CCl4 (0.2–3.500 M) at
CCl4 CBrCl3, 1-C10H21Br (bromodecane), C6H11Br 60 °C by 3:2 mixtures of [H2W10O32]4– and [HW10O32]4–
(bromocyclohexane), and t-C4H9Br by reduced (het- (3.9–33 mM) were studied in detail under pseudo.
eropoly blue) forms of the decatungstate anion first-order conditions (excess CCl4). The precise
[W10O32]4– 226–230 and of CCl4 by the reduced form of stoichiometry of the reaction was complicated by the
the Keggin anion α-[PW12O40]3– 231,232 (fully oxidized presence of two reducing agents.
forms shown) was investigated by Sattari and Detailed study of the reduction reaction could
Hill.233,234 The “Q” salts (Q = tetra-n-butylammo- easily have been simplified by use of [HPW12O40]3–,
nium cation) of the decatungstate and undecatung- a well-defined 1-e– heteropoly blue. Nonetheless, the
stophosphate anions (i.e., Q4[W10O32] and α-Q 3 - decatungstate mixture was chosen deliberately. Use
[PW12O40]), dissolved in dry degassed acetonitrile, of the reduced decatungstate mixture, prepared pho-
were reduced photochemically by irradiation in the tochemically, better represented a key thermal step
presence of 2-butanol (sacrificial reductant) under an in the catalytic photochemical reduction of haloal-
inert atmosphere. 235 Oxidative titration of the re- kanes by [W10O32]4–, a much more versatile236–240 and
duced anions with Ce(IV) gave 1.0 ± 0.05 electrons active photocatalyst than [PW12O40]3–.
156 Chemical Reviews, 1998, Vol. 98, No. 1 Weinstock

Although products of both 1- and 2-e– reductions Savinov252 and co-workers published a full paper on
of haloalkanes were observed (Table 8), the relatively the reduction of H+ to H2 by stoichiometrically
high and exclusive yield of CHCl3 (entry 1, Table 8) reduced heteropolytungstates. Shortly thereafter,
was indicative of a 2-e–-reduction stoichiometry (eq Darwent noted that when the photocatalytic chem-
151). istry described by Papaconstantinou was carried out
in the absence of O2, H+ was reduced to H2 by reduced
POM anion intermediates.253 Under standard condi-
tions, the electrochemical potential for reduction of
By monitoring the rate of formation of fully oxi- O2 to H2O by reduced POM anions is +1.23 V more
dized [W10O32]4– the following rate law was deter- exergonic than that for reduction of H+ to H2.
mined: Nonetheless, in many cases the latter reaction is
spontaneous.
More thorough investigations of the chemical pro-
cesses outlined in the above-mentioned reports were
published by Papaconstantinou254 and Darwent255 in
1985. In the same year, Hill236 outlined the broad
log–log plots of rate vs [CCl4] and vs [(H2W10O32)4– scope of POM-catalyzed photooxidations and photo-
+ (HW10O32)4–] had slopes of 1.02 (R = 0.99) and 0.93 dehydrogenations of organic compounds. In a 1993
(R = 0.97), respectively. The observed rate law, review,248 Hill and Prosser-McCartha commented
product distributions, and relative reaction rates that “most of the scope and mechanistic details in
argued for the existence of at least short-lived radical polyoxometalate photochemistry have been clarified
intermediates in the reactions listed in Table 8. since 1985 and thirty research groups have now
Comparison with other studies of reductive deha- published nearly two hundred papers in this area”.
logenation reactions argued that rate-limiting elec- Despite this activity, and the use of both aerobic and
tron transfer was more likely than halogen atom anaerobic reaction conditions, little detailed kinetic
abstraction. The remaining question was whether data pertaining specifically to the homogeneous (no
the electron-transfer step was dissociative (i.e., colloidal Pt catalyst236,253,254 present) reduction of H+
whether addition of an electron to the σ∗ antibonding by reduced POM anions has been published.
LUMO of the alkyl halide occurred with concerted Savinov252 deoxygenated aqueous solutions of
carbon-halogen bond rupture) or whether short-lived [SiW12O40]4– and used zinc amalgams to produce
but discrete anion radicals were the first products of mixtures of the protonated 1- and 2-e–-reduced
electron transfer. It was concluded that while ad- heteropoly blues, H[SiW12O40]4– and H2[SiW12O40]4–,
equate experimental evidence could be cited in sup- respectively, at low pH (0.1 N H2SO4. After separat-
port of the existence of anion radicals resulting from ing the heteropoly blue solutions from the amal-
the reduction of aromatic halides and from the gams, progress of reaction was followed using three
reduction of polyhalogenated halides, such as CBr4 techniques: UV-vis and ESR spectroscopy and de-
or CCl4,225 the situation was not simple in the case termination of evolved H2g. At 22 °C the 2-e–-reduced
of common aliphatic halocarbons. anions reduced H+ to H2 according to the overall
stoichiometry:
VII. Reduction of Inorganic Oxidants by Reduced
POM Anions
A. Introduction The reaction rate was constant over a dramatic
The use of inorganic electron acceptors, such as range of H+ concentrations (0.0001–6 N H2SO4). Also,
Br2,241 permanganate,242,243 and inorganic per- rather than second order in [H2(SiW12O40)4–], as the
oxides,244–247 is often encountered in POM synthesis. reaction stoichiometry might have implied, the pro-
Taking advantage of the oxidative stability of many cess obeyed first-order kinetics (eq 154, kobs = 1.2 ±
POM anions, researchers have also used ozone.120 10–3 s–1). Analogous results were obtained by UV–
Though some reactions, such as the reduction of
H+ 248 and nitrite,249 have been studied in the context
of photo- and electrocatalyzis, relatively few kinetic vis spectroscopy and by ESR observation of the
studies concerning POM reductions of common inor- accumulation of paramagnetic H[SiW12O40]4–.
ganic oxidants have been reported. These observations suggested that the reaction
Because of its complexity, and its important role might occur via slow unimolecular decay of
in POM-catalyzed oxidations of organic and other H2[SiW12O40)4–, followed by rapid comproportion-
substrates, the reduction of dioxygen (O2) by reduced
ation:
POM anions will be discussed separately (section
VIII).

B. Hydrogen Cations
In 1981 Papaconstantinou submitted a communi-
cation to the Royal Society of Chemistry outlining the
photocatalytic oxidation of organic compounds by
POM anions and oxygen.250,251 At the same time, This hypothesis was confirmed by adding fully oxi-
Electron-Transfer Reactions of POMs Chemical Reviews, 1998, Vol. 98, No. 1 157

dized [SiW12O40]4– to the reaction: Although less H2 Despite the difference in charge between the X =
was produced, kobs remained unchanged. Si or Ge and the X = P or As anions, the insensitivity
Like Savinov, Papaconstantinou254 observed that of kobs to their formal reduction potentials (Table 9)
decreases in optical density of acidic (pH 1) solutions is strong evidence that the reaction with NO2– (pH
of mixtures of 1- and 2-e– photochemically reduced 5) occurred via an inner-sphere mechanism (cf.
[SiW12O40]4– anions obeyed pseudo-first-order kinet- Figure 5). While still acknowledging that at lower
ics. pH values outer-sphere 1-e– reduction of HNO2 by
Although reduction of H+ by H[SiW12O40]4– was not [XFeII(OH2)W11O39] n – was plausible, the authors fa-
observed by Savinov, Darwent253 noted that at high vored an inner-sphere pathway involving coordina-
[H+], the reaction was thermodynamically favorable tion of HNO2 to Fe(II) (eqs 158–161). This reaction
and in the presence of colloidal Pt, kinetically facile. scheme is consistent with the labile nature of octa-
Darwent later studied the rate of evolution of H2 hedral Fe(II) aquo complexes (cf. reduction of O2 by
during irradiation (λ > 300 nm) of acidic (0.5 M H2- less labile Ru(II) aquo complexes; section VIII.E).
SO4) solutions containing [SiW12O40]4–, CH3OH, and
colloidal Pt254 Discussion of these and related248
heterogeneous photocatalytic reactions is beyond the
scope of this review.

C. Nitrite
The reduction of nitrite (NIIIO2–) to nitric oxide
(NIIO) by aqueous [XFeII(OH2)W11O39]n – anions (X =
Si or Ge, n = 6; X = P or As, n = 5) was examined in
the context of a larger effort to develop efficient
catalysts for multielectron electrochemical reduction
of nitrite to ammonia.269 Kinetic runs were carried
out at constant ionic strength (µ = 0.1 M) under
pseudo-first-order conditions of excess nitrite. The
reaction was first-order with respect to the total
nitrite concentration and with respect to [XFeII-
(OH2)W11O39) n –]. Observed second-order rate con- Three lines of electrochemical evidence were pre-
stants were proportional to [H+] at pH values above sented in support of this reaction scheme: (1) At pH
the pKa of nitrous acid (HNO2, pK a = 3.3) 5 in acetate buffer, the reduced anions formed by
and independent of [H+] at lower pH values. The electrochemical reduction of the lacunary anions,
largest rate constant was observed below the pK a of [XW11O39]( n +3)–, which although strongly reducing,
nitrous acid, indicating that HNO2 was the ac- lacked a suitable coordination site for NO2–, did not
tive form of the oxidant. The initially observed reduce nitrite at an appreciable rate. (2) Because of
nitrogen-containing product of the reaction was the the affinity of Fe(III) for hydroxide anion, the outer-
nitrosyl adduct [XFeII(NO)W11O39]n –, and 2 equiv of sphere reduction of HNO2 to give NO and uncoordi-
[XFe II(OH2)W11O39]n – were consumed in its forma- nated HO– (eq 162) would possess a significantly
tion: smaller driving force than the inner-sphere electron-
transfer step shown in eq 159. This argument

assumes that an outer-sphere electron-transfer step


The identity of the adduct was confirmed by inde- would generate free hydroxyl anion (eq 162). Al-
pendent synthesis from [XFeII(OH2)W11O39]n – and though outer-sphere electron transfer to HNO2 might
NO. In addition, at pH 5 the authors observed a be expected to give HNO2–, the latter is a high
correlation between the formal potentials of the energy molecule whose generation would likely re-
XFeIIIW11O39]n–/[XFeIIW11O39]( n +1)– couples (X = Si, quire a greater driving force than that provided by
Ge, P, and As) and the rate constants for nitrosyl the FeIII/II–heteropolyanion couple.256 (3) At pH 5
complex, [XFeII(NO)W11O39]n –, formation (Table 9). no reduction of NO2– was observed upon electro-
Table 9. Correlation of Formal Potentials (Ef) for
chemical reduction of [XFeIII(OH2)W11O39](n –1)– to
[XFeIIIW11O39] n –/[XFeIIW11O39]( n +1)– Couples with Their [XFeII(OH2)W1lO39] n –. Yet at more negative poten-
Kinetic Properties tials (reduction of the tungsten oxide framework),
catalytic currents corresponding to the reduction of
coordinated NO to NH3 were observed. No reduction
of NO was observed upon reduction of the iron-free
lacunary anions at pH 5. These observations implied
the accessibility, at the more negative electrode
potentials, of an alternate route for the formation of
[XFeII(NO)W11O39] n –. It was proposed that coordi-
158 Chemical Reviews, 1998, Vol. 98, No. 1 Weinstock

nated HNO2, in [XFeII(O2NH)W11O39ln –, was reduced At issue is the value of the inner-sphere reorgani-
not by electrons originating in Fe(II), but rather by zation energy, λin. Theoretically calculated values
more energetic electrons available at the more nega- have ranged from λin = 78 to λin, = 108 kcal mol–1.
tive electrode potentials. Combined with an estimated outer-sphere reorgani-
The third argument is less straightforward than zation value of λout = 36 kcal mol–1, total reorganiza-
the first two lines of evidence. The observation of tion energies of λtot = 114–144 kcal mol–1 are
electrocatalytic reduction of NO to NH3 exclusively obtained. At the same time, from the rate of hetero-
in the presence of iron implies a role for Fe(II) geneous electron transfer to NO2 from a Pt electrode,
coordination of NO. However, the lack of electro- a value of λin = 79 kcal mol–1 was calculated. To
catalytic NO reduction in the absence of iron does obtain an accurate value of the reorganization energy
not rule out the possibility that electrostatic factors of the NO2+/NO2 couple, Lund and Eberson (1997)
might have been responsible for the lack of reactivity measured the rate of homogeneous-phase outer-
of NO2– with reduced lacunary anions, which are sphere electron transfer from [CoIIW12O40]6– to NO2+
more highly negatively charged than reduced Fe(II)- (eq 166).257
substituted analogues.
Finally, it should be noted that any electron-
transfer reaction that proceeds along an outer-sphere
pathway does not require (and cannot benefit from)
a redox catalyst that operates via an outer-sphere
pathway. The reason is that the same outer-sphere The reactions of Bu4N+ and mixed Bu4N+–K+ salts
electron-transfer reaction should be just as facile at of [CoIIW12O40]6– with NO2+BF4– were studied in
an electrode surface held at the potential of the acetonitrile and in nitromethane at 26 °C. Although
(catalyst)ox/(catalyst)red couple.256 While true for el- due consideration was paid to the extent of ion
ementary electron-transfer steps, this reasoning need pairing in solution and to the reduction potentials of
not preclude the occurrence of intermediate outer- [CoIIIW12O40]5– and NO2+ under reaction conditions,
sphere electron-transfer reactions in multistep reduc- considerable uncertainty in these parameters was
tion pathways. acknowledged. For the Bu4N+ salts, bimolecular rate
In the electrocatalytic reduction of NO2– to NH3, constants of kobs = 0.097 ± 0.013 M–1 s–1 and 12.5 ±
[XFe II (NO)W 11 O 39 ] n – was further reduced at the 1.5 M–1 s–1 were measured in acetonitile and ni-
cathode. The authors provided strong circumstantial tromethane, respectively. In acetonitrile, a rate
evidence that the electrons accepted by the NO ligand constant of kobs = 0.073 M–1 s–1 was measured for
coordinated to Fe(II) were delivered via the tungsten the K+ salt.
oxide framework of the polyanion rather than directly Although confronted with large uncertainties in
from the electrode surface to the bound NO ligand.249 key parameters, Marcus theory (eqs 1 and 4) was
used to calculate reorganization energies for the
D. Nitronium Ion reaction (eq 166) from the experimentally determined
rate constants. Analysis of the data obtained in
It is generally accepted that electrophilic aromatic nitromethane was complicated by unacceptably large
nitration by the nitronium ion (eq 163) proceeds via uncertainty in the reduction potential of NO2+ in this
bond formation between NvO2+ and ArH to give a solvent. From the rate constant for reaction of the
nitrocyclohexadienyl cation followed by loss of a Bu4N+ salt of [CoIIW12O40]6– in acetonitrile (k obs =
proton. 257 0.097 ± 0.013 M–1 s–1), a total reorganization energy
(λ tot) of from 80 to 85 kcal mol–1 was calculated.
The reorganization energy associated with the
NO2+/NO2 couple was then calculated using λtot = ( λ1 1
A possible alternative mechanism, outer-sphere + λ22 (eq 43) where λ11 and λ22 are total reorgani-
electron transfer from ArH to NO2+ (eq 164) followed zation energies associated with the CoIII/IIW125–/6– and
by the reactions shown in eq 165, has been ruled out NO2+/NO2 couples, respectively (section II.B.4). Tak-
on the basis of theoretical and experimental evidence- ing λ22 equal to 140 kcal mol–1, a reorganization
energy of λ11 = 25 kcal mol–1 for the CoIIIW125–/6–
couple gave a total reorganization energy, λtot = (25
+ 140)/2 = 82.5 kcal mol–1. This value was consis-
tent with the λtot value of between 80 and 85 kcal
mol–1 obtained from kobs using eqs 1 and 4. This
agreement was presented as evidence in support of
a high value (~140 kcal mol–1) for the reorganization
A primary theoretical difficulty with the outer- energy of the NO2+/NO2 couple.
sphere mechanism is that the reorganization energy At the same time, assuming a λ22 value of 140 kcal
of the NO2+/NO2 couple, which would have a large mol–1 for the NO2+/NO2 couple, good fit of the ko b s
influence on the reaction rate, is probably prohibi- value (kobs = 0.073 M–1 s–1) obtained for reaction of
tively large. The large reorganization energy is a the mixed Bu4N+–K+ salt of [CoIIW12O40]6– in aceto-
consequence of the change in geometry, from linear nitrile with eqs 1 and 4 required that a value of λ 1 1
NO2+ to bent NO2, in the transition state. The actual = ~40 kcal mol–1 be assigned to the CoIII/IIW125–/6–
value of the reorganization energy, however, has been couple. This higher value for λ11–a consequence of
controversial. the somewhat lower kobs value obtained for the mixed
Electron-Transfer Reaction of POMs Chemical Reviews, 1998, Vol. 98, No. 1 159

Bu4N+–K+ salt vs that obtained using the Bu4N+ any [H+] dependence reported by Banerjee: By using
salt-was tentatively attributed to a greater degree Na2SO4 to (conscientiously) maintain a constant ionic
of ion pairing between K+ and [CoIIW12O40]6–. strength while varying [H2SO4], Banerjee had inad-
In light of the uncertainty that appears to exist in vertently buffered the reaction solutions at the pK a
the value of λ11 (section III.C), these conclusions must of bisulfate. After controlling for the pH-leveling
be regarded, at least for the present, as tentative. In effect of SO42–, Mehrotra observed the [H+] depen-
particular, assignment of a high reorganization en- dence Banerjee had apparently anticipated.
ergy to the NO2+/NO2 couple requires that the reor- Recognizing that lacunary POM anions could be
ganization energy associated with the CoIII/IIW125–/6– used to stabilize high oxidation states of coordinated
couple (λ11) be small (i.e., ~25 kcal mol–1). In section metal ions, Shilov261,262 used S2O82– to oxidize lan-
III.C, a concern was raised that this value for λ 1 1 thanide- (M = Pr and Tb) and actinide- (M = Am)
might be considerably underestimated. To the extent substituted heteropolytungstates, [MIIIP2W17O61]7– to
that the true value for λ11 might be greater than 25 [MIVP2W17O61]6–. The oxidation of [AmIIIP2W17O61]7–
kcal mol–1, the conclusions reached here by Lund and to [AMIVP2W17O61]6– was first order in S2O82– and
Eberson would be compromised. zeroth order in polyanion. More recently You and
At the time of this writing, Professor Eberson has Gu263 reproduced Shilov’s results using [TbIII-
indicated his intention of cross-checking the NO2+/ (PW11O39)2]11– complexes.
NO2 reorganization energy by using other sub- Using a similar synthetic strategy, Pope247 recently
strates.89 More generally, however, given the con- prepared Mn(IV)-substituted heteropolytungstates,
tinued and frequent use of the cobaltotungstate [MnIV(OH)XW11O39] n – (X = Zn, B, and Si), by reaction
anions as quantitative probes for investigating outer- of the corresponding Mn(II) complexes with S2O82–.
sphere electron-transfer reactions-the present re- This result might help explain some puzzling kinetic
cently published work of Lund and Eberson being a data, reported several years earlier by Mumtaz and
case in point-accurate redetermination of the reor- Zubairi.264 They found that the 1• e– oxidation of
ganization energy of the CoIII/IIW125–/6– couple appears [MnIISiW11O]6– by S2O82– was one-half order in
justified. S2O82–, and although details of the data analysis are
sketchy, they appear to have observed a first-order
E. Peroxodisulfate dependence of reaction rate on [(MnIISiW11O39)6–].
They proposed a mechanism involving rapid and
Banerjee investigated the 2-e– reduction of per- reversible homolytic dissociation of S2O82– followed
oxodisulfate by [CoIIW12O40]6– under pseudo-first- by slow oxidation of Mn(II) to Mn(III) by SO4• –. This
order conditions (excess [S2O8]2–):258,259 proposal should be reevaluated.

F. Chlorate and Periodate


Ayoko, Iyun, and co-workers265 investigated the
reduction of ClO3– to Cl– by [CoIIW12O40]6– in aqueous
At 50 °C the reaction was first order in [S2O82–] (10– HClO4 solution at an ionic strength of µ = 1.0 M
80 mM) but zeroth order in [(CoIIW126–)] (0.1–0.2 (NaClO4) (eq 171).
mM), and the rate was independent of [H+] (0.001-
0.50 M; ionic strength kept constant by the addition
of Na2SO4) and ionic strength (µ = 1.56–2.06 M,
KNO3). The experimental rate law (eq 168) was

Uncatalyzed, the reaction was very slow even at


60 °C. In the presence of Ag+, however, more rapid
rates were seen and the following rate law (eq 172)
consistent with rate-limiting unimolecular dissocia- observed. Even in the presence of Ag+, however, the
tion of S2O82– (eqs 169 and 170).

reaction was not particularly fast: At 60 °C k o and


k1 were reported to be (1.28 ± 0.10) × 10–4 M–1 s–1
and (2.22 × 0.20) × 10–4 M–3 s–1, respectively.
Ayoko also used [CoIIW12O40]6– to investigate the
Banerjee noted that previous investigators had reductions of BrO3– 265 and IO4–.259
observed an [H+] dependence that was not evident Banerjee258 investigated the rate of 2-e– reduction
here. However, Banerjee’s value of k1 ((0.73 ± 0.03) of periodate, IVIIO4– (1.0–60.0 mM total [I(VII)]), to
× 10–6 s–1) agreed with the literature value (1.0 × iodate, IVO3–, by [CoIIW12O40]6–. The reaction was
10–6 s–1), as did the observed activation parameters studied over a pH range of 0.7 to 3.0 ([H+] = 1.0–
(∆ H ‡ = 30 kcal mol–1 and ∆S ‡ = 6.9 cal K–1 mol–1). 200 mM, HNO3) at µ = 0.10 M (KNO3). In aqueous-
Investigating the same reaction, Mehrotra260 made acid solution, periodic acid exists in equilibrium with
an interesting observation regarding the absence of its hydrated form, orthoperiodic acid, H5IVIIO6 (eq
160 Chemical Reviews, 1998, Vol. 98, No. 1 Weinstock

173). At 25 °C periodic acid dominates (K = 3.4 ×


10–2). The acid mixture has an apparent pK a of 1.6.

where M = Li, Na, or K. Values for k', determined


At 60 °C the reaction stoichiometry was dependent at low [H+] (0.01 M) (i.e., higher [MnO4–), followed
upon the [CoIIW126–]/[I(VII)]T ratio, where [I(VII)]T is the order k' Li < k'Na < k'K. The following elementary
the total concentration of I(VII). When the [CoIIW126–]/ reaction steps were proposed (subsequent reduction
[I(VII)]T ratio was 1:10 (excess I(VII)), the stoichi- steps were assumed to occur rapidly):
ometry was 0.9 ± 0.02 CoIIW126–:I(VII); at a ratio of
6:1 (excess CoIIW126–), the stoichiometry was 2.01 ±
0.03.
At pH 1 under pseudo-first-order conditions (excess
I(VI), 60 °C), the reaction was first order in
[CoIIW126–] At constant [CoIIW126–] (0.2 mM), the
reaction was first order in [I(VII)]T at low [I(VII)]T
(~1–3 mM) and gradually became independent of
[I(VII)]T at higher concentrations (>15 mM I(VII); i.e., It was noted that the order of rate enhancement
saturation kinetics). A plot of kobs vs [H+] (0.001– by the alkali-metal cations was in keeping with the
0.200 M, pH 3–0.7) possessed a nearly zero intercept greater degrees of ion pairing frequently observed for
and increased asymptotically to a limiting value at the larger cations. The authors recognized, however,
high [H+]. This suggested that the acids, HIO4 and that ion pairing (eq 178) shifted the potential of the
H5IO6, rather than their conjugate bases, were the CoIIIW125–/CoIIW126– couple to more positive values.52
reactive species. At constant [Na+] at pH 1, increase
in the ionic strength had no effect on rate. However,
increases in Na+ or K+ concentrations decreased the
reaction rate, and the inhibitory effect of Na+ was Thus, ion pairing (eq 178) alone would have been
slightly greater than that of K+. expected to decrease the reaction rate, with k1Li > k1 N a
Proposed mechanistic schemes included the acid > k1K, opposite to the order actually observed. Faced
dissociation of H5IO6 and association between with these enigmatic results, the authors proposed
[CoIIW12O40]6– and H5IO6 prior to electron transfer that the greater polarizability of the larger cations
(relative reactivities Of HIO4, vs HIO6 were not might make them more kinetically facile “bridges”
explicitly addressed). The observation of both as- for electron transfer. Recent data50 appear to argue
sociation-complex formation and inhibition by specific against this proposal (see section II.C).
cations, M+ (as opposed to the specific-cation catalysis An alternative explanation is that the larger cat-
observed in oxidations by [COIIIW12O40]5–), led the ions better facilitated the formation of ternary com-
authors to propose an “inner-sphere” mechanism plexes, [(MnO4–)(M+)(CoIIW12O406–)]6– (i.e., larger
involving an intermediate complex, formed by “bond- formation constants for the ternary aggregates).
ing of a periodate oxygen to a surface tungsten atom”. Analogous considerations pertain to the effects of
(The inhibitory effect of alkali-metal cation was alkali-metal cation pairing (eq 178) on the electro-
attributed to competition with H5IO6 for association chemical potential52 of the CoII/IIW12O405–/6– couple.
with [CoIIW12O40]6–.) It is puzzling why the observa- The Marcus cross-relation (eq 17) was used to
tion of association-complex formation and inhibition evaluate the reduction of protonated HMnO4 by
by alkali-metal cations should require recourse to setting the [H+]-dependent rate constant, ko (eq 175),
such a mechanism. The observations are adequately equal to Kk1. Division of ko (6.0 M–2 s–1) by the
accounted for by ion pairing between [CoIIW12O40]6– literature value for K (2.99 × 10–3 M–1) gave k1 =
and alkali-metal cations, M+ (section II.C),52 and the 2.01 × 103 M–1 s–1. Application of the Marcus cross-
formation of donor–M+–acceptor association com- relation to the evaluation of k1 assumed the absence
plexes prior to electron transfer (section II.D). of association-complex formation prior to electron
transfer. This appears justified: Values of k' (eq 175)
G. Permanganate were small relative to k1 (eq 177) (i.e., ternary
complex formation did not appear to be important in
Ayoko recently studied the reduction of perman- the reactions of [Mn CoIIW12O40](6–n)– ion pairs with
ganate by [CoIIW12O40]6– at 25 °C (eq 174) by moni- neutral HMnO4).
In the Marcus calculation, a literature value of k2 2
for the HMnO4/HMnO4– self-exchange reaction (mea-
sured at µ = 1.0 M, NaClO4) was used. For the
CoIII/IIW125–/6– self-exchange reaction, the published
k11 value5 determined at µ = 0.1 M (LiCl) was used
toring the decrease in absorbance of MnO4– (525 nm) without extrapolation to prevalent ionic strength (see
under pseudo-first-order conditions (lo-fold excess of section III.A above). Nevertheless, the Marcus cross-
CoIIW125–) at high ionic strength (µ = 1.0 M, NaClO4) relation gave a value of k1cal (5.37 × 103 M–1 s–1) in
over a range of H+ concentrations (0.01–0.50 M, relatively good agreement with experimental obser-
HClO4).267 The following rate law was observed: 3 –1 –1
vation, k1 e x p = 2.01 × 10 M s .
Electron-Transfer Reactions of POMs Chemical Reviews, 1998, vol. 98, No. 1 161

In light of section III.C, it should be noted that tation) of superoxide to hydrogen peroxide and diox-
although use of the Marcus cross-relation (eq 17) ygen. The reaction (eq 183) is spontaneous over the
requires knowledge of self-exchange rate constants, entire pH range in water and goes to completion even
in this case k11 for the CoIII/IIW125–/6– couple, no at pH 14.268 At neutral pH, KpH 7 = 4 × 1020.
assumptions regarding the reorganization energy
inherent in k11 need be made.

VIII. Reduction of Dioxygen by Reduced POM The bimolecular disproportionation reaction is


Anions promoted by protons; it is relatively rapid at pH
values well below 4.69 (the pK a of superoxide radical,
A. introduction HO 2• ) and quite slow at pH values above 4.69 (eqs
POM anions can serve as electron-transfer agents 184 and 185).269 (The pKa of H2O2 is 11.75, well above
in the selective homogeneous thermal or photochemi- that of HO2• .)270
cal oxidations of organic or other substrates by
dioxygen.113,144 The catalytic cycle is often repre-
sented by two steps: (1) selective oxidation of a
substrate (SubH2, eq 179) by the oxidized form of a
POM anion (POM,ox) and (2) reoxidation of the
reduced POM anion (POM,red) by dioxygen (eq 180).

The pH dependences of the reduction potentials of


additional reactions involved in the stepwise 4-e–
reduction of O2 to water or hydroxide anions have
been reviewed.271
The two steps can occur consecutively, as shown, or
simultaneously; the POM anions can be reduced by C. The O 2 /O 2• – Self-Exchange Reaction
one electron each, as shown, or by more than one
electron. The selective oxidation of substrate (eq 179) The value of the self-exchange rate constant, k1 1
may be either thermal or photochemical (i.e., utilizing (and implicitly, reorganization energy, λ, for reduc-
photoexcited POMox*). However, the reduction of tion of O2 to O2• – is central to theoretical evaluation
dioxygen by POMred anions, the topic of this section, of rate constants for reactions that involve outer-
is generally a thermal process. (The kinetics and sphere reductions of O2. However, as recently as a
mechanisms of a number of homogeneous POM- decade ago, a great deal of uncertainty surrounded
catalyzed oxidations by O2 are discussed by Kozhevni- experimentally determined rate constants for the O2/
kov114 in this issue of Chemical Reviews.) O2• – self-exchange reaction.272 The reason for this
uncertainty was that estimated k11 values, obtained
B. Single-Electron Reduction of O 2 by Marcus treatment of rate constants measured for
cross-reactions of O2 or O2• – with inorganic electron-
Reduced POM anions reduce dioxygen by both donor or electron-acceptor complexes, ranged from
inner- and outer-sphere mechanisms. In the outer- 10–7 to 107 M–1 s–1. Questions pertinent to explana-
sphere reduction of dioxygen, the first step is electron tion of this variation included273 (1) was the Marcus–
transfer to O2: Hush theory applicable to systems involving small
molecules in which the size, charge, and extent of
solvation of the oxidized and reduced species (e.g.,
O2 and [O2• (H2O)n ]–) differed greatly from one an-
In water at low pH values, the reduction is ac- other, and (2) were some of the reactions used to
companied by protonation of the superoxide anion to determine k11 values truly adiabatic outer-sphere
the superoxide radical (HO2• –; eq 182), which has a electron-transfer processes?
pK a of 4.69.268 In 1988 Zahir, Espenson, and Bakac273 found that
the relationship between kinetic and thermodynamic
data for single-electron reductions of O2 by a series
of Cr(II) complexes obeyed that predicted by Marcus.
At pH 0 in water, the single-electron-reduction Moreover, examination of published data showed that
potential, E°, of O2 is +0.12 V vs NHE. This value the highly scattered k11 values previously reported
is based on unit concentrations of O2 and its reduction had all been derived from cross-reactions involving
products in water. If unit pressure is taken as the reduction by O2• – anions (the italics are the authors’),
standard state of O2, the more often quoted value of while k11 values determined from reductions of O2
–0.05 V vs NHE applies.269 The standard reduction were in good agreement with one another. They
potential of O2 decreases with pH. At pH values suggested that some complicated interactions were
substantially above the pK a of HOO• (e.g., above ~pH likely involved in the reductions of oxidants by O2•–
7), E° = –0.16 V for unit concentrations of O2 in anions. From their and others’ rate data for reduc-
water (–0.33 V for 1 atm O2). tions of O2 by substitution-inert inorganic com-
Also pertinent to discussion of O2 reduction reac- plexes, k11 was determined to lie in the range 1-10
tions is the disproportionation (or in biology, dismu- M–1 s–1.
162 Chemical Reviews, 1998, Vol. 98, No. 1 Weinstock

Table 10. Polyanion Reduction Potentials, Relative Quantum Yields for the Photochemical Formation of
Reduced Polyanione, and Rate Constants for Electron Transfer from the Reduced Polyanione to O2

In a 1989 study by Lind, Shen, Merényi, and Jons- × 10–4 M) water at pH 1 (HClO4). The pseudo-first-
son,274 the rate constant for the O2O2• – self-exchange order rate constant for reoxidation of [PW12O40]4– by
reaction in water was measured directly under O2 (k = 2.2 × 10–2 s–1, excess O2) was measured by
alkaline conditions (pH > pK a of HOO• ) by means of monitoring the rate of decrease of the absorption
18
O/16O isotope marking. A second-order rate con- band (751 nm) attributed to the reduced polyanion.
stant of k11 = 450 ± 160 M–1 s–1 was found. From The same technique was used by Akid and Dar-
this value, a reorganization energy of λ = λin + λo u t went255 several years later to determine the rates of
= 45.5 ± 0.7 kcal mol–1 was calculated from the oxidation of a series of reduced Keggin anions by O2.
Marcus relationship (eq 1). The authors pointed out Microsecond flash photolysis of aqueous solutions
that in outer-sphere reactions in which the effective containing the Keggin heteropolytungstate anions
radii of the reacting partners differ substantially, λ o u t [XW 1 2 O 4 0 ]n– (POMox X = P(V), Si(IV), Fe(III), 2H+,
values are not strictly additive. Consequently, k1 1 and Co(II); 2 × 10–6 M) and i-PrOH (1.7 M) in
values calculated from cross-reaction rate data (such oxygenated water ([O2] = 2 × 10–4 M) at pH 1 (H2-
as those determined by Zahir, Espenson, and Bakac) SO4) gave the 1-e–-reduced anions, [XW12O40)(n + 1 ) –
might easily be a few orders of magnitude smaller ((5–8) × 10–7 M). The rate of reduction of O2 to HOO•
than 450 M–1 s–1, the directly determined value. was measured by monitoring the rate of decay of the
In 1993 Merényi, Lind, and Jonsson275 demon- absorbance attributed to the heteropoly blue anion
strated that application of the Marcus equation (eq [XW12O40]( n +1)– (eq 186). The reaction was first order
1, using λ as an adjustable parameter) to kinetic and
thermodynamic data for the oxidations of 25 closed-
shell (singlet ground-electronic-state) organic com-
pounds by 3O2 gave an excellent fit. The best fit gave
λ = 37.4 ± 2.5 kcal mol–1, which corresponded to an
apparent k11 of 2 M–1 s–1. Notably, the closed-shell in [POMred], [O2], and [H+], which suggested a rate
organic compounds studied were similar in size to expression of the form
the substitution-inert Cr(II) and Ru(II) complexes
(atomic radii ranging from 3.4 to 6.8 Å) evaluated by
Zahir, Espenson, and Bakac.273 The atomic radii of
most POM anions also fall within this range.
Reductions of O2 by three general categories of
POM anions have been reported: (1) isopoly- and The first-order dependence of rate on [H+] sug-
heteropolytungstates, (2) d-electron-containing first- gested that HOO•, rather than OO• –, was involved
row transition-metal- (and ruthenium-) substituted in the rate-determining step. More direct evidence
heteropolytungstates, and (3) mixed-addenda vana- for the formation of HOO•, which at low pH values
domolybdophosphates. disproportionates rapidly to O2 and H2O2 (eq 184),
or of H2O2, was not reported. Although the rate of
D. Reduction of O2 by Reduced Isopoly- and oxidation of the reduced Keggin anions by H2O2 was
Heteropolytungstates also not discussed, it has been shown277 that the rate
of reaction of a closely related heteropolytungstate,
In 1980 Papaconstantinou276 and co-workers re- the 1-e–-reduced Wells-Dawson anion [P2W18O62]6–
ported that various isopoly- and heteropolyanions, (POMox, shown) with hydrogen peroxide is an order
when irradiated in the presence of organic com- of magnitude slower than the rate of oxidation of the
pounds, were reduced to the corresponding blues, and reduced anion by O2.
the organic compounds present were oxidized. Two POM reduction potentials and termolecular rate
years later he reported250 that when coupled with constants for the reduction of O2 measured at con-
reoxidation of the reduced POM anions by O2, the stant O2 and H+ concentrations ([O2] = 2 × 10–4 M,
photochemical reactions could provide the basis for pH = 1) are shown in Table 10.
photocatalytic oxidations of organic substrates by Plot of ln k vs –∆G° /RT (eq 8 and Figure 8) gave
dioxygen (eqs 179 and 180). As part of a preliminary a slope of 0.52, very close to the value of 0.5 predicted
survey of POM reactivity, he used microsecond flash by Marcus (eq 1; |∆G°'| ^ λ). In keeping with the
photolysis of [PW12O40]3–, in the presence of 2-pro- first-order dependence of rate on [H+] noted above,
panol, to generate the 1-e–-reduced heteropoly blue the [H+]-dependent value for the reduction potential
anion [PW12O40]4– in oxygen-saturated ([O2] = ~2.5 of O2 was used to calculate ∆G° (see footnote in Table
constants followed the reduction potentials of the
anions, the more thermodynamically favored reoxi-
dation reactions occurring more rapidly: [P2W18O62]7–
< [SiW12O40]5– < [H2W12O40]7–. Although the antici-
pated reduction product, H2O2, was not observed (hy-
pathetical concentrations were too low in any event),
the expected stoichiometry, 2 equiv of 1-e–-reduced
POM anion per equivalent of O2, was established by
following the consumption of dioxygen (eq 188).

Figure 8. Plot of ln k vs –∆G°/RT for the reduction of O2 At pH 1 the rate of reaction of [P2W18O62]7– with
by reduced heteropolytungstate anions. O2 (kobs/[O2] = 4.8 M–1 s–1) was 24 times faster than
its rate of reaction with H2O2 (k obs/[H2O2] = 0.2 M–1
10). The y intercept of the plot in Figure 8 (ln Z – s–1). In addition, in contrast to results reported for
λ/4RT) is 8.9, from which a collision frequency of Z the reduction of O2 by 2- and 4-e–-reduced (Pr4N)3–
= 1011 M–1 s–1 gives λ = 38.9 kcal mol–1. [PW12O40] salts in acetonitrile,278 Hiskia and Papa-
If the reorganization energies for the O2/O2– (λ 1 1 ) constantinou found that the rates of reoxidation of
and POMn / (n +1)– (λ 22) self-exchange reactions were 2-e–-reduced anions in water were greater than those
strictly additive (i.e., λ = 0.5(λ 11 + λ22)), a value of λ of 1-e–-reduced anions.
= 0.5(45.5 + 22.3) = 33.9 kcal mol–1 would have been At constant pH and ionic strength, the reoxidations
expected on the basis of published data for λ 11274 and of 1-e–-reduced heteropolyanions were first order in
λ22.23 However, large differences in size and shape [POMred] and in [O2]. However, in plots of k obs vs pH,
between O2 (linear) and spherical electron donors some reactions (reoxidation of [P2W18O62]7– and
have been shown to cause deviations from strict [SiW12O40]5–) appeared independent of [H+] at low
additivity. For example, radius ratios, r2 2 /r 11 (effec- and high [H+] values and first order in [H+] at
tive radius of O2, r11 = ~2.5–3.0 Å), of ~2.5 are intermediate [H+] values. At low pH values of ~1,
expected to give experimental cross-reaction λ values pseudo-first-order rate constants of kobs = 3.8 × 10–3,
that are ~6 kcal mol–1 greater than those predicted 1.0, and > 5.5 s–1 were observed, and at high pH
on the basis of strict additivity.275 Due to the values (>3), kobs values of 0.0, 2.7 × 10–2, and 0.15
discrepancy in size and shape between O2 and the s–1 were observed for the reactions of O2 with
Keggin anion (relatively spherical, r = 5.6 Å), similar [P2W18O62]7–, [SiW12O40]5–, and [H2W12O40]7–, respec-
deviation from strict additivity should be anticipated. tively. (The kobs values observed at pH 1 are com-
The result obtained from the combined data of parable to those reported in Table 10.) The reduced
Papaconstantinou250 and Akid and Darwent,255 (λ e x p anions [P2W18O62]7– and [SiW12O40]5– remained un-
– λcalc) = 38.7–33.9 = 5.0 kcal mol–1, is in line with protonated over the pH range studied. By contrast,
this expectation. kobs values for reoxidation of [H2W12O40]7–, which is
The rate of oxidation of tetra-n-propylammonium protonated at pH values below 4.9, increased to a
(Pr4N+) salts of a mixture of 2- and 4-e–-reduced maximum as the pH was increased from near 0 to 1
[PW12O40]3– anions in air-saturated acetonitrile was and then decreased, reaching a near-constant value
followed by visible absorption spectroscopy at 25 at pH values greater than 3.
°C.278 In acetonitrile-containing tetrabutylammo- Because two of the anions were known to remain
nium perchlorate (0.1 M) supporting electrolyte, the unprotonated over the pH range studied, ion pairing
half-wave potential for, the first reduction of (Pr4N)3- between the POM and H+ ions was excluded as an
[PW12O40] was –0.1 V vs NHE compared to +0.218 explanation for the pH dependence of kobs. One
V at pH 1 in water. Plot of ln(absorbance) vs time alternative possibility considered was that the pH
for reoxidation of the reduced POM anion mixture dependence resulted from protonation of a dioxygen
gave a straight line, which implied nearly identical molecule coordinated to the reduced POM anion. The
rate constants for the reactions of O2 with the formation of an adduct between the 3O2 diradical and
different reduced POM species present. Using the an unpaired electron in the b2 (dx y ) orbital of a
known solubility of O2 in acetonitrile, an aggregate tungsten atom in the 1-e–-reduced POM anion was
pseudo-unimolecular rate constant of 3.7 × 10–2 s–1 suggested. Given the kinetic stability of the W-O-W
was calculated for the reoxidation reactions. Analo- bonds in most heteropolytungstates, a 7-coordinate
gous pseudo-unimolecular rate constants for reoxi- tungsten complex (eq 189), although sterically con-
dation of [PW12O40]4– in water at pH 1 were 2.2 ×
10–2 s–1 250 and 15 × 10–3 s–1.225
Hiska and Papaconstantinou studied the rate
of reoxidation of 1-e–-reduced heteropolyanions
(P2W18O62]7–,+ [SiW12O40]5–, and [H2W12O+40]7–) as +a
function of [H ] by adjusting the ratio of [H ] to [Na ]
at high ionic strength (0.5 M, ClO4–).277,279 As
anticipated the magnitudes of pseudo-first-order rate gested, was deemed more likely than an intermediate
164 Chemical Reviews, 1998, Vol. 98, No. 1 Weinstock

Table 11. Standard One-Electron-Reduction Potentials of Mn(III)-Substituted α-Keggin Heteropolytungstate


Anions

involving coordination of O2 to either a terminal mined that the rates of formation of labeled H2O2 and
(M=O) or bridging (M–O–M) oxide (O2–) anion of the water were much greater than the rates of exchange
reduced polyoxoanion. of these products with oxygen atoms in the polyoxo-
On the basis of this reasoning, the inflection points metalate ([W10O32]4– and [P2W18O62]6–) frameworks.
seen in the plots of kobs vs pH might have cor- Hence, it was unequivocably established that no
responded to the pK a value of the coordinated dioxy- oxygen incorporation into the polyoxoanions occurred
gen molecule and/or might have signified a mecha- during, or immediately after, electron transfer to O2.
nism involving rate-limiting electrophilic attack of H+ Duncan and Hill argued that if inner-sphere com-
on coordinated dioxygen. Attempts to observe O2- plexes, such as that shown in eq 189, were involved,
bound intermediates by ESR, as had been observed at least some of the 17O-labeled oxygen would have
in the reactions of Mn(II)-substituted heteropolyan- been incorporated into the polyanion frameworks. At
ions with O2 in anhydrous apolar organic sol- the same time, the authors acknowledged that be-
vents,178,179 were unsuccessful, and the issue was left cause the oxygen-exchange chemistry of hypothetical
unresolved. POM-O2 adducts could not be assessed, the absence
In retrospect, however, these data need not be of O2 incorporation did not entirely rule out an inner-
interpreted as evidence of an inner-sphere mecha- sphere mechanism. However, these 17O-labeling
nism. The effects of pH on kobs might be attributed studies, combined with the close correlation with
not to [H+], but rather to the changes in Na+ Marcus theory reported by Akid and Darwent,255
concentrations required to maintain the high ionic point strongly to an outer-sphere mechanism for the
strength (µ = 0.5 M) as H+ concentrations were reduction of O2 by aqueous solutions of reduced
varied. At pH values near 0, Na+ concentrations isopoly and heteropolytungstate anions.
would be very small (~0.003 M), while at pH values
greater than 3, Na+ concentrations would exceed E. Reduction of O2 by d-Electron-Containing,
0.499 M. At these high Na+ concentrations, signifi- Transition-Metal-Substituted Heteropolytungstates
cant ion pairing can be expected. The effect of ion
pairing is to shift the reduction potentials of the POM In 1970, building on the work of Baker et al.285 and
anions to more positive values, making ∆G° for Malik and Weakley,244,246 Tourné and Weakley242
electron transfer to O2 less favorable. Hence, the prepared a series of isostructural Mn(II)-substituted
inflection points observed in the plots of pH vs ko b s α-Keggin heteropolytungstate anions, [Xx MnII(OH2)-
might reflect pKf values for Na+-ion pairing (Kf = W 11 O 39 ] ( 1 0 –x )–, where X = Zn(II), B(III), Si(IV), Ge(IV),
formation constant for ion pairing) rather than pKa and P(V). In addition, Mn(III) analogues were pre-
values for protonation of coordinated dioxygen. At pared by chemical oxidation of Mn(II), and the reduc-
the same time, the pH dependence of the reduction tion potentials of the [Xx MnIII/II(OH2)W11O39](9– x ) – / ( 1 0 – x ) –
potential of the O2/O2– couple (i.e., the effect of this couples were determined (Table 11).
potential on ∆G°) must still be considered. In only one case, X = B, was the reduction potential
A general correlation between POM anion reduc- of the Mnlll/Mnll couple sufficiently negative that
tion potentials and rates of reoxidation by O2 was also oxidation of Mn(II) by molecular oxygen was kineti-
observed by Kuznetsova and co-workers,280,281 who cally, as well as thermodynamically, feasible. Thus
studied the oxidation of 0.05 M solutions of reduced [BMnIII(OH2)W11O39]6– was prepared by exposing
mixed-addenda vanadotungstophosphate anions, slightly acidic aqueous solutions of [BMnII(OH2)-
[PVn W 1 2 –n O 40 ] (3+n )– (POMox forms indicated), in 40% W11O39]7– to air for several days at room temperature.
acetic acid. They found the at moderate tempera- Inner-sphere oxidation of Mn(II) to Mn(III) in
tures neither [PVIVW11O40]5– nor [PVIVVVW10O40]6– coordinatively unsaturated Mn(II)-substituted het-
reacted with O2. However, at pH 3 reduced eropolytungstates in dry apolar solvents was first
[PV3W9O40]6– anions7– (POMox shown) reacted slowly, reported by Katsoulis and Pope in 1984.178,179 The
while [PV4W8O40] anions reduced to the same organic-soluble polyanions were prepared via ex-
extent reacted rapidly. Although the degrees of ion change of alkali-metal countercations by organic-
(H+ or Na+) pairing present in each case were not soluble quaternary alkylammonium cations, such as
evaluated, more negatively charged anions generally tetra-n-heptylammonium bromide.286 The heteropoly-
possess more negative reduction potentials.32 anions [XMnII(OH2)W11IIO39]n – (X = P(V), Si(IV), Ge(IV),
Duncan and Hill282 provided additional support for and B(III)) and α2-[P2Mn (OH2)W17O61]8– were trans-
an outer-sphere mechanism in the reactions of re- ferred into nonpolar solvents (benzene and toluene)
duced isopoly- and heteropolytungstates with O2. by shaking aqueous solutions of the POM anions with
They used 17O NMR to investigate the temporal equal volumes of organic-solvent solutions of the
evolution of 17O in products resulting from the quaternary ammonium salts.
reactions of reduced POM anions with 17O2.283,284 Following separation and drying of the POM anion
Upon reduction of 17O2 by POMred, the 17O label was solutions, coordinatively unsaturated Mn(II) com-
initially observed in solution as H217O2 and H217O. plexes (e.g., [SiMnIIW11O39]6–) were observed by ESR
From a series of control experiments, it was deter- spectroscopy. These coordinatively unsaturated com-
Electron-Transfer Reactions of POMs Chemical Reviews, 1998, Vol. 98, No. 1 165

plexes formed adducts with a variety of donors, such autocatalytic: Order of reaction with respect to
as alcohols, pyridine, and halides, and in several [(PFeII(OH2)W11O39)5–] was ~1.6–1.8. The high,
cases, reversible formation of O2 adducts was ob- nonintegral concentration dependence suggested that
served at low temperatures. (Even at –70 °C, how- additional equivalents of [(PFeII(OH2)W11O39)5–] were
ever, exposure of solutions of [BMnIIW11O39]7– to O2 being oxidized by products resulting from the reduc-
resulted in instantaneous formation of [BMnIII- tive activation of O2.
W11O39]6– Formation of the O2 adducts (eq 190) was A closely related and analogously complex system
reversed by bubbling N2 through the solution; warm- was the topic of a detailed kinetic study performed
ing the oxygenated solutions to ambient temperature by Toth and co-workers.291 They found that the rate
resulted in irreversible oxidation of Mn(II) to Mn- constant governing the electrocatslytic reduction of
(III). (By contrast, even at 150 °C and 100 psig O2, H2O2 by aqueous solutions of [SiFeIII(OH2)W11O39]5–
oxidation of K6[SiMnII(OH2)W11O39] in water is slow.)120 was 9 × 102 M–1 s–1. At high [SiFeIII(OH2)W11O39]5–
concentrations, the expected stoichiometry for the
reduction of H2O2 to water (two electrons per equiva-
lent of H2O2, eqs 191 and 192) was observed. How-

In the presence of polar substances, such as water,


pyridine, methanol, acetone, acetonitile, or chloride
ions, no reaction was observed upon oxygenation.
This behavior was attributed to competition of the ever, under catalytic conditions (substrate-to-catalyst
donor molecules for the available (sixth) coordination ratios greater than unity), fewer than two electrons
site on Mn(II) and was interpreted as evidence that were consumed for each H2O2 molecule consumed. A
the reaction with oxygen occurred via direct inner- catalytic mechanism involving an Fe(IV) intermedi-
sphere coordination of O2 to Mn(II). ate (or an HO• adduct of the Fe(III) complex) pro
Paradoxically, however, no reaction was observed duced by reaction of [SiFeIII(OH2)W11O39]5– with
when the phase-transfer-derived POM anion solu- hydroxyl radicals (HO•) was proposed. The noninte-
tions were scrupulously dehydrated; it was suggested gral stoichiometry observed in the electrocatalytic
that the oxygen adducts might be stabilized by hydro- reduction of H2O2 was attributed to competition
gen bonding to a water molecule. Efforts to fully between two pathways, each involving further reac-
characterize an O2 adduct were only partially tion of the Fe(IV) intermediate: reaction of the Fe(IV)
successful: Addition of the spin trap 5,5-dimethyl- intermediate with itself to give [SiFeIII(OH2)W11O39]5–
1-pyrroline N-oxide (dmpo) resulted in solutions that and an equivalent of H2O2 (overall stoichiometry two
had ESR spectra consistent with the formation of electrons per equivalent of H2O2 consumed) or reduc-
Mn–O2–dmpo adducts and that over time gave rise tion of the Fe(IV) intermediate by H2O2 to give
to ESR signals tentatively interpreted as resulting [SiFeIII(OH2)W11O39]5– and HOO• , and via dispropor-
from the oxidation of dmpo by superoxide. In addi- tionation of HOO• , O2 and H2O2 (0.67e– per equiva-
tion, in solutions of oxygenated [SiMnIIW11O39]6–, lent of H2O2 consumed). The mechanistic proposal
hindered phenols were oxidized to cyclohexadienyl was supported by detailed kinetic analysis of pro-
radicals. However, no oxidation was observed when posed elementary reactions: [SiFeII(OH2)W11O39]5– +
the phenols were added to oxygen-free solutions of H2O2, [SiFeIII(OH2)W11O39]5– + HO• , and [SiFeIII-
[SiMnIIIW11O39]7–. It was thus suggested that attack (OH2)W11O39]5– + O2• –.
by a ‘Mn(III)O2•” moiety or related polyanion–O2 Significantly, the rate constant for the reaction
species was responsible for single-electron oxidation between [SiFeII(OH2)W11O39]5– and H2O2 (9 × 102
of the phenols. M–1 s–1) was only 1 order of magnitude greater
Analogous reversible O2–adduct formation and than the value reported for the reaction of H2O2
inner-sphere oxidation of Mn(II) to Mn(III) and of Co- with [Fe(H2O)6]2+, even though the reaction with
(II) to Co(III) was observed upon bubbling of Oz. into [SiFeII(OH2)W11O39]5– was more favorable by 680 mV.
anhydrous toluene solutions of tetraalkylammonium It was noted that this insensitivity to thermodynamic
salts of Mn(II)- and Co(H)-substituted Keggin and driving force argued against an outer-sphere mech-
Wells-Dawson anions.287 anism.
Reactions of oxygen, superoxide anions, hydrogen In related work Bart and Anson studied the l-e-
peroxide, and hydroxyl radicals with aqueous solu- oxidation of the second-transition-series cogener,
tions of Fe(II)-substituted heteropolytungstate anions [PRuII(OH2)W11O39]5–,755,293,294 by O2 in water.295,296
were studied in connection with the catalytic oxida- Rong and Pope,75 who observed clean metalation of
tion of H2S to S° by [PFeII(OH2)W11O39]5– and O2 (eq the [PW11O39]7– “lacunary” ligand upon addition of
172, SubH2 = H2S)211,288–290 and the electrocatalytic [RuII(H2O)6]2+, found that warm aqueous solutions of
reduction of H2O2 by [SiFeIII(OH2)W11O39]5–.291,292 At the resultant product, K5[PRuII(OH2)W11O39] (~5
20 °C aqueous [PFeII(OH2)W11O39]5– was oxidized mM), were oxidized to K4[PRuIII(OH2)W11O39] in 0.5
slowly to [PFeIII(OH2)W11O39]5– by O2. The reaction h by bubbling with O2.
rate (k = 3.3 × 10–4 M min–1 L–1)289 was independent The reduction potential of the [PRuIII/II(OH2-
of [H+] over a pH range of 2 to 5. The reaction was W11O39]4–/5– couple (+0.21 V vs Standard Hydrogen
166 Chemical Reviews, 1998, Vol. 98, No. 1 Weinstock

Table 12.II Reaction Conditions and Rate Consants (Measured and Calculated) for the Reduction of O2 by
Cs5[PRu (OH2)W11O39] and Other POM Salts in Water

Electrode, SHE) was similar to that of the FeIII/FeII cussed in section VII.C. This error may have con-
couple of [PFeIII(OH2)W11O39]6– (+0.28 vs SHE).75 tributed to overestimation of kcalc (Table 12).
By contrast, oxidation of aqueous solutions of Closer examination of the data in Table 12 points
[PMnII(OH2)W11O39]6– by O2 (the potential of the out a problem rarely discussed in applications of
[PMnIII/II(OH2)W11O39]4–/5– couple is +1.12 V vs SHE) Marcus theory, that is, how to deal with pH-depend-
has not been reported. In addition, the pK a associ- ent driving forces. The ∆G° values listed in Table
ated with deprotonation of the aquo ligand coordi- 12 were estimated by the author of the present
nated to Ru(II) in [PRuII(OH2)W11O39]5– (eq 193) was review using the pH-dependent chemical potentials
5.1. The potential of the [PRuIII/II(OH2)W11O39]4–/5– associated with the 1-e– oxidations of the POM
anions and the 1-e– reduction of O2. However, in
calculating kcalc values, Bart and Anson assumed that
electron transfer from [PRuII(OH2)W11O39]5– to O2
occurred as shown in eq 194, that is, prior to pH-
couple was independent of [H+] between pH 0 and 4 dependent deprotonation of [PRuIII(OH2)W11O39]5– or
and decreased by ~80 mV (to ~+0.13 V) as the pH protonation of O2–. 256 Thus, despite the large varia-
was increased from 4 to 6.
Rong and Pope presented several lines of evidence
that suggested that n-electron density from the d
orbitals of the Ru(II) (d6) ion was partially delocalized tions in ∆G° values associated with the overall
onto the polytungstate ligand.75 This previously reactions, a single theoretical kcalc value (10 kcal
unobserved phenomenon, which could influence elec- mol–1) was obtained. The validity of Bart and An-
tron-transfer reactions of Ru-substituted polytung- son’s approach is suggested by the near constancy
state complexes, was attributed to the more extended in reported kexp values and is consistent with the
d-orbital structures of the second- and third-row above-noted lack of a dependence of rate on pH.
transition-metal cations. At the same time, the pKa for deprotonation of
Bart and Anson used spectroscopic and electro- [PRuII(OH2)W11O39]5– to [PRuII(OH2)W11O39]6– is 5.1
chemical methods to measure the rate of room- (eq 193). Thus, determination of kcalc for the reaction
temperature oxidation of cesium salts of [PRuII- carried out at pH 5.75 (entry 3 in Table 12) would
(OH2)W11O39]5– by O2 at several ionic strengths and appear to have required more explicit consideration
over a range of pH values that encompassed the pK a of the 80 mV larger driving force associated with the
of the Ru-coordinated aquo ligand (Table 12).295,296 elementary reaction step shown in eq 195. Curiously,
In all cases changes in absorption attributed to the however, the rate constant observed at pH 2 (entry
[PRuII(OH2)W11O39]5– anion corresponded to the 2-e– 1, Table 12) was not substantially smaller than those
reduction of O2 to H2O2; no nonintegral stoichiom- measured above the pKa of [PRuII(OH2)W11O39]5–
etries were observed. In addition, the reaction was (entries 2 and 3, Table 12).
first order in [PRuII(OH2)W11O39]5– and in O2. No
dependence of rate on pH (between pH 2 and 5.4) was
observed.
Using the Marcus cross-relation (eq 17), k22 = 1 ×
107 M–1 s–1 for electron self-exchange between An additional line of experimental evidence in sup-
[PRuII(OH2)W11O39]5– and [PRuIII(OH2)W11O39]4–, and port of an outer-sphere mechanism was also presen-
k11 = 450 M–1 s–1 for the O2/O2– self-exchange ted:75,295 When polar σ-donor and π-acid ligands were
274
reaction, the authors calculated a single theoretical added to aqueous solutions of [PRuII(OH2)W11O39]5–
rate constant of 10 M–1 s–1 for all three reaction the rates of exchange of the aqua ligand on Ru(II)
conditions (Table 12). The relatively good agreement were several orders of magnitude smaller than the
with kexp values was regarded as evidence for outer- rates of oxidation of [PRuII(OH2)W11O39]5– by O2.
sphere electron transfer.
The inapplicability of the k11 value determined F. Reduction of O2 by Reduced
directly from the 18O2/16O2– self-exchange reaction Vanadomolybdophosphates
(450 M–1 s–1) to reactions of O2 with much larger Reductions of O2 by solutions of reduced (heteropoly
electron donors has been noted274,275 and was dis- blue) vanadomolybdophosphate anions are reviewed
Electron-transfer Reactions of POMs Chemical Reviews, 1998, Vol. 98, No. 1 167
114
by Kozhevnikov in this issue of Chemical Reviews. Dr. Rajai H. Atalla, Dr. Carl J. Houtman, Dr. Cindy
In general a number of factors have made full E. Sullivan, Richard S. Reiner, and Matthew J.
characterization of the mechanisms of reoxidation of Birchmeier; their probing and intellectually demand-
reduced vanadomolybdophosphate anions by O2 an ing questions, over the course of an ongoing effort to
extraordinarily difficult task. First, POM anions develop POM-based catalyst systems for the selective
consisting largely of molybdenum addendum atoms oxidation of agricultural products by oxygen, led the
are more labile than their tungsten analogues. As a author to address the issues covered in this article.
result, equilibrium mixtures of polyanions of varying The topic of this article was also influenced by the
composition are often rapidly established in aqueous work of Dr. P. Banejee and co-workers, who pub-
solution. For example, the most widely cited prepa- lished a similarly titled article in 1993. In addition,
ration of H5[PV2Mo10O40]117 yields a product whose warm thanks are extended to Professor L. C. W.
elemental analysis is consistent with the proposed Baker for sharing his thoughts regarding the struc-
formula. However, 51V and 31P NMR spectra of ture and reactivity of Keggin dodecatungstocobaltate
aqueous solutions of the product reveal an equilib- anions, which he originally prepared almost 50 years
rium mixture of acids, H3+n [PVn Mo 12–n O40] (n = 1,2, ago, and for his insight regarding the work of
and 3), in which H5[PV2Mo10O40] is the major prod- Rasmussen and Brubaker, who measured the rate
uct.86,119,120 Also, when more than one vanadium of electron self-exchange between these complexes
atom is present (e.g., [PVn Mo 12–n O 40 ] (3+ n )–, n > 1), over 30 years ago. In addition, thoughtful feedback
mixtures of positional isomers are present.121 Third, on draft copies of the manuscript by Professors Ivan
dissociation of V(V) and V(IV) from both fully oxidized V. Kozhevnikov, Fred C. Anson, Craig L. Hill, and
and reduced (V(IV)-containing) anions, respectively, Lennart Eberson is gratefully noted, while special
has been observed. The extent of dissociation varies thanks are due to Professor Eberson for his gracious
with the number of vanadium addendum atoms in assistance in the critical reevaluation of published
the POM structure, pH, the extent of reduction of the values for key theoretical parameters. Finally, the
POM anion, and more than likely (although rarely editorial assistance of Janel Showalter of the Forest
mentioned), concentration.188,189,297–300 Despite con- Products Laboratory is warmly acknowledged. The
siderable effort, these complications have hindered cover art was created by Jennifer Cowan.
detailed characterization of the mechanisms of O2
reduction by heteropoly blue vauadomolybdophos- X. References
phate complexes.
Nonetheless, Matveev, Kozhevnikov, Kuznetsova,
Neumann, and their co-workers, who have studied
these reactions in the greatest detail, propose inner-
sphere mechanisms involving intermediate com-
plexes between O2 and V(IV) addendum atoms of the
POM anion.85,116,145,301–306 In only one report have
data, which might be regarded as evidence for an
outer-sphere mechanism, been presented282 (see dis-
cussion of this report by Kozhevnikov).114
According to the prevalent view, an open coordina-
tion site on the V(IV) ion is provided by the breaking
of a V(IV)–O bond between the reduced (d1) vana-
dium ion and a bridging p-oxygen atom (i.e., within
a V–O–MO linkage) of the parent heteropoly blue, a
process consistent with the labile nature of the
vanadomolybdophosphate anions.

IX. Acknowledgments

The topic of this review, written while the author


was a visiting scientist in the laboratory of Professor
Craig L. Hill at Emory University, arose from ques-
tions encountered over the course of a long-standing
collaboration between the USDA Forest Service,
Forest Products Laboratory, Madison, WI (the au-
thor’s home institution), and the Emory University
Department of Chemistry. The hospitality, encour-
agement, and intellectual vision of Professor Hill are
warmly acknowledged, as are the hospitality and
assistance of co-workers at Emory, in particular,
Marc Theune, Dr. Qin Chen, Dr. Michael Wemple,
Jeff Rhule, Jennifer Cowan, and Kun Liu. Acknowl-
edgment is also due the author’s colleagues at the
USDA Forest Service, Forest Products Laboratory:
168 Chemical Reviews, 1998, Vol. 98, No. 1 Weinstock
EIectron-Transfer Reactions of POMs Chemical Reviews, 1999, Vol. 98, No. 1 169
170 Chemical Reviews, 1998, Vol. 98, No. 1 Weinstock

Das könnte Ihnen auch gefallen