You are on page 1of 13

Journal of The Electrochemical Society, 149 4 S21-S33 2002

S21

0013-4651/2002/1494/S21/13/$7.00 The Electrochemical Society, Inc.

A Century of Organic Electrochemistry


Henning Lund*
Department of Chemistry, University of Aarhus, DK 8000 Aarhus C, Denmark
2002 The Electrochemical Society. DOI: 10.1149/1.1462037
Available electronically March 12, 2002.

The electrochemistry of organic compounds in the 20th century


is built on the work done in the 19th century. The invention of the
battery, the Volta pile, in 1800,1 was essential for the development of electrolysis as such experiments require a current running
during an extended period. Volta, however, did not regard the chemical reactions observed at the electrodes as essential for the function
of the battery and his position in science was so high that no preparative electrolyses were made until Faradays experiments in the
eighteen thirties. Faraday was the first who made an electroorganic
synthesis2 by electrolyzing an acetate solution and obtaining a gaseous product, ethane. The anodic oxidation of salts of fatty acids to
hydrocarbons with loss of carbon dioxide was developed by Kolbe3
to become the first useful electroorganic synthesis, and it was later
in the century extended by Brown and Walker4 to electrolysis of
salts of monoesters of dibasic acids resulting in dimerization to esters of dibasic acids. The first electrochemical reduction of an organic compound seems to be the reductive dehalogenation of
trichloromethanesulfonic acid to methanesulfonic acid5 at a zinc
electrode.
During the second half of the 19th century preparative organic
electrochemistry bloomed and a very optimistic view on the synthetic possibilities of the new method both for laboratory and technical applications was expressed. In this classical period several
oxidations and oxidative substitutions as well as reductions of nitro
compounds, carbonyl derivatives and dehalogenation reactions were
performed; however, in most cases mixtures of products were obtained. Many of these investigations were made by the groups of
Elbs, Gattermann, Haber, Lob, and Tafel. Their publications appeared mostly in Ber. Dtsch. Chem. Ges., Z. Elektrochemie, and J.
Prakt. Chem.
In 1898 Haber6 published a classical paper on the stepwise reduction of nitro compounds; in this he realized that by using a constant current density the effective reduction potential would gradually become more negative and that it was essential for a selective
reaction to keep the potential at the working electrode constant.
However, control of the potential could only be made manually and
there was no easy way to find the optimal potential for a given
reaction.
20th Century
The organic electrochemistry of the 20th century may be divided
into three periods: from 1900 to 1940, from 1940 to 1960, and since
1960.
1900 to 1940.The situation in the first period was not much
different from the last decade of the 19th century with regard to
technique, but the optimism had almost disappeared during the period and only few publications appeared on organic electrochemistry
between 1910 and 1940. A few groups continued doing research,
mainly those of F. Fichter, S. Swann, and H. D. Law. Surveys of the
work done in that period may be found in the book by Fichter7 and
the two reviews by Swann.8
Only a few electroorganic reactions became industrial processes,
and some of the more important ones were indirect reactions where
the electrochemical step was a regeneration of an inorganic reagent.
As early as 1900 it was realized9 that in many cases it was an

* Electrochemical Society Active Member.

advantage to separate the electrochemical step which was the reoxidation of CrIII to CrVI from the chemical oxidation in which
CrVI was used to oxidize naphthalene or anthracene to the quinones. Another process in which this method was used was the bleaching of raw montan wax, and this reaction has been running commercially for more than 70 years by Hoechst now Clariant on a 10,000
ton-per-year t/yr scale. Montan wax is extracted from lignite and
consists of esters of long-chain acids and alcohols; after hydrolysis
of the esters the alcohols are oxidized by CrVI to the acids.
A production of sorbitol and mannitol by cathodic reduction of
glucose was established by Atlas Powder company10 on a 1400 t/yr
scale, but some years later the electrolytic reaction was replaced by
high-pressure catalytic hydrogenation.
1940-1960.In that period electroanalytical methods, such as
polarography and voltammetry at solid electrodes, were applied to
the study of many organic molecules, and the results used for analysis and in a few cases for guidance of electrolysis at controlled
potential. Also the use of aprotic media was begun, both for cathodic
and anodic reactions.
Polarography.Polarography, current-voltage curves obtained at
the dropping mercury electrode DME, was invented by
Heyrovsky11 and was in the beginning mostly used for analysis of
inorganic compounds; however, a publication on the polarography
of nitrobenzene appeared early.12 The classical photographically recording polarograph was constructed by Heyrovsky and
Shikata.13
Between 1925 and 1940 some organic compounds such as
fructose,14 aromatic and halogenated carbonyl compounds,15
acetylacetone,16 and cystine17,18 were investigated by polarography.
The two books in 1941 on polarography by Kolthoff and
Lingane19 and the new edition of Heyrovskys book20 in German
had a great effect on polarographic research. Examples of compounds investigated polarographically in the nineteen forties are
phenylsubstituted olefins and polynuclear aromatic hydrocarbons,21
aliphatic and aromatic halogen compounds,22 nitro compounds,23
diazonium salts,24 and N-heterocycles.25,26
A seminal paper by Lingane27 demonstrated that the potentials
found by polarography could be used for selective reductions at a
controlled potential at a macroelectrode, and that preparative electrolysis at the potential of the limiting polarographic current could
establish the electrode reaction in polarography. Lingane demonstrated that 9-2-iodophenylacridine could selectively be reduced to
the 9-2-iodophenyldihydroacridine without reducing the carboniodine bond. The introduction of potentiostats28,29 made it easier to
perform electrochemical reactions at controlled potential.

In the nineteen fifties and nineteen sixties polarographic investigations of many types of organic compounds were made and the

S22

Journal of The Electrochemical Society, 149 4 S21-S33 2002

results used for analysis and/or for a guidance for controlled potential electrolysis, mainly reductions at a macro-mercury electrode.
Several books were published on polarography30-34 and in Analytical Chemistry Wawzonek wrote a survey of polarographic publications on organic electrochemistry every second year. Later in the
century other electroanalytical techniques became more popular, although still valuable polarographic investigations were made, foremost by Zuman.
A theory for fast polarography, oscilloscopic polarography, at a
dropping mercury electrode was developed by Matsuda
and Ayabe;35 the electrode reactions were divided into reversible,
quasi-reversible and irreversible according to the rate constant of the
heterogeneous electron transfer. The problems and their solutions
are similar to those encountered in linear sweep LSV and cyclic
voltammetry CV.
Anodic voltammetry in aqueous or aqueous-alcoholic media is
limited by the available potential range. The use of
acetonitile/NaClO4 as medium was introduced by Lund36 in investigations on the electrochemical oxidation of alcohols and aromatic

hydrocarbons. Preparative oxidations at controlled potential were


made, sometimes in the presence of pyridine as base/nucleophile.
The conclusion was that the oxidation consisted of a loss of electrons from the aromatic system which then, by loss of protons or
reaction with a nucleophile, formed the products. This was in contrast to the prevailing ideas that anodic oxidations proceeded
through electrolytically generated reagents, such as peroxides.
Industrial processes 1940-1960.The most important industrial
application of organic electrolysis introduced in the period 19401960 was the electrochemical fluorination ECF of organic substrates, and the electrochemical manufacture of fluorinated compounds has since continued to be an important industrial process.
The oldest one, the Simons ECF37 is an electrolysis of the organic
compound in anhydrous hydrogen fluoride aHF at a nickel anode,
and it results in a perfluorination of the substrate. All hydrogens in
the compound are replaced by fluorine, double bonds are saturated,
some functional groups are retained, but a certain degradation of
longer chains occurs.
The mechanism of the fluorination has been discussed during the
years; the involvement of radical cations,38 fluoride ions, elemental
fluorine, fluorine radicals,39 and high-valence nickel fluorides40 have
been suggested and the importance of a layer of partly fluorinated
polymers at the anode has been underlined. Recently it has been
proposed that the fluorination goes through NiIV fluoride which in a
two-electron three-center bonds binds to the substrate. The NiIV
may be reduced in a process in which the adsorbed substrate is
oxidatively fluorinated or, alternatively, the substrate may be oxidized without fluorination to give carbocations and reaction products
from these or give tar; the reactive NiIV is continuously regenerated
at the anode.41
The Phillips Petroleum Company has developed an electrochemical fluorination using an electrolyte of KF2HF with a porous
carbon anode. The process is thought to involve electrolytic generation of elemental fluorine, and the fluorination is believed to take
place within the pores of the anode. Fluorination of hydrocarbons
gives partly and fully fluorinated products. The method is a useful
complement to the Simon process.
Later, other electrochemical partial fluorination reactions were
developed. In these reactions the substrate is oxidized to a radical
cation or cation which reacts with fluoride ion, so it is essential to
perform the oxidation in a medium in which the nucleophilic reac-

tivity of the fluoride ion is not diminished too much by solvation.


The preferred media are R3 N nHF or R4 N nHF, 42 but other
media have also been used.43 A review on electrochemical partial
fluorination has recently appeared.44
Another industrial production, although on a much lower scale,
also began operating about 1950, the anodic methoxylation of furan
and furan derivatives to 2,5-dihydro-2,5-dimethoxyfurans.45 Such
compounds are derivatives of 1,4-dicarbonylbutene and as such are
readily transformed to other compounds. A number of heterocyclic
compounds were prepared from 1,4-dicarbonyl-2-butenes. The reaction was discontinued after some years but later reintroduced on a
larger scale by BASF and Otsuka Chemical

1960-2001.After about 1960 organic electrochemistry underwent a rapid development. A characteristic feature of the period was
that new electroanalytical techniques made it possible to investigate
the mechanism of the electrode reactions more efficiently than previously. Another development is the introduction of indirect electrolysis using organic and organometallic mediators and a growing
interest of the electrochemistry of bioorganic systems. Furthermore,
new industrial applications of organic electrochemistry appeared. A
few areas of the development can be mentioned here, but can only
be sketchily described.
Electroanalytical techniques.In the years after 1960 the further
development of electroanalytical methods made it possible to obtain
a more detailed understanding of the different steps in the electrochemical reactions. Both the theory for different types of reactions
and the instrumentation were developed, and the introduction of
simulation of the various kinds of electroanalytical signals was a
great step forward.46,47 A commercially available simulation program, Digisim, is now available.46 A monograph on the theory of the
electroanalytical techniques has been published,48 and discussions of
the methods for reaction studies has recently appeared.49-51
One of the most widely used techniques is cyclic voltammetry;
publications dealing with the influence of the rate of the heterogeneous electron transfer and the rate of follow-up reactions on the
shape of the curves were published by Saveant and Vianello52 and
by Nicholson and Shain.53 Sometimes it is an advantage to include
the whole curve in the interpretation by calculation of the convolution integral which transforms the voltammogram into a
polarogram-like curve.54,55
When the electron transfer is followed by a rapid follow-up reaction, a high sweep-rate is needed to outrun the reaction, and for
microelectrodes of conventional diameter about 1 mm the voltage
drop caused by the ohmic resistance is difficult to compensate electronically. By the use of ultramicroelectrodes diam 2-10 m higher
sweep-rates may be employed.56 Using suitable ohmic-drop compensation it is possible to make an ohmic-drop-free electrochemical
investigation in the megavolt per second range which corresponds to
the development of a diffusion layer having only a few nanometers
thickness.57-59 However, unfortunately the ratio between the capacitative and Faradaic current is proportional to the square root of the
sweep rate so at very high sweep rates the Faradaic current is just a
small bump on the large capacitative current. At high sweep rates
the rate of the heterogeneous electron transfer may set a limit for the
use of ultramicroelectrodes.
Another application of ultramicroelectrodes is in the scanning
electrochemical microscope SECM. It consists of a microelectrode
radius 1-25 nm, with the distance between the tip and the substrate
being piezo-controlled. At a distance between the tip and the con-

Journal of The Electrochemical Society, 149 4 S21-S33 2002


ductive substrate similar to the radius of the tip, a molecule reduced
at the tip may diffuse to the substrate and be reoxidized, before it
reacts in other ways; this gives an increase in the current as the
apparent concentration is increased. If a fast follow-up reaction occurs, the changes from the expected current at different distances
give information on the rate of the follow-up reaction. The experiments have some analogies to rotating ring-disc measurements. Using SECM the rates of fast reactions such as the dimerization of
acrylonitrile in DMF60 and the rate of the oxidative dimerization of
4-nitrophenolate61 were determined.
The SECM allows the observation of the electrochemical behavior of a single molecule by trapping a small volume of a dilute
solution of the electroactive species, e.g., a ferrocene derivative,
between the tip and the conductive substrate. An anodic current was
observed when the molecule moved in or out of the electrodesubstrate gap.62
The SECM has also been used to investigate the reverse electron transfer, in which the electron is transferred from the reduced
form to the oxidized in an uphill reaction. Such thermodynamically unfavored reactions are common in living systems, and the
driving force here is believed to be a proton electrochemical potential gradient across a membrane. A model for such a membrane is
two immiscible solutions, and the role of the proton to produce a
potential drop across the interface is taken over by a supporting ion
common to both solutions. Using ferrocyanide and 7,7,8,8tetracyanoquinodimethane TCNQ, the forward and back rate constants for electron transfer between water and dichloromethane have
been measured.63,64
A SECM may be used to probe the redox activity of individual
living cells and to measure the rate of transmembrane charge transfer. Significant differences were detected in the redox response
given by normal human breast ephithelial cells and metastatic breast
cancer cells.65 Intracellular redox potentials and concentrations of
redox components in the cell can be evaluated noninvasively.66
Oxidative bursts of reactive oxygen species ROS are used
by cells to fight bacteria and viruses; such oxidative bursts produced
by human fibroblasts can be followed electrochemically by using a
carbon fiber platinized ultramicroelectrode placed within a few microns of a single living cell, when the membrane is punctured with
a micrometer-sized sealed pipette. In spite of experimental difficulties, it has been possible to obtain electric signals indicating that the
oxidative bursts consist of a mixture of cytotoxic chemicals: H2 O2 ,

NO, ONO
2 , and NO2 . The cell survived the puncture and could
perform a similar burst after a couple of hours.67
Media, electrodes.Protic solvents have continued to be used; a
newly employed solvent is 1,1,1,3,3,3-hexafluoropropan-2-ol,
CF3 CHOHCF3 . 68 The solvent has an unusually low nucleophilicity,
a high ionizing power and a high hydrogen bonding strength which
makes a it good solvent for cation radicals.
Although aprotic media had been employed before 1960 both for
reductions69 and oxidations,36 the use of such media for investigations of electrode reactions and synthesis became much more common after that date. For oxidations acetonitrile with sodium perchlorate as supporting electrolyte and a Ag/Ag-reference electrode was
first employed.36 Later methylene chloride,70 sometimes with added
trifluoroacetic acid and its anhydride to remove nucleophiles,71 was
used when acetonitrile was too nucleophilic and reacted with the
intermediates. Liquid sulfur dioxide offers a solvent with very low
nucleophilicity,72 but the limited solubility of supporting electrolyte
allows preparative electrolysis in only a few cases.73
For reductions N,N-dimethylformamide DMF, acetonitrile
MeCN and dimethylsulfoxide DMSO have been used. Acetonitrile is slightly more acidic than DMF and DMSO, but less likely to
lose hydrogen by atom abstraction. Acidic impurities may be removed by activated aluminum oxide.74 When hydrogen atom abstraction is a problem, as in SRN1 reactions, liquid ammonia can be
used.75 Methylamine with LiCl as supporting electrolyte has been
the preferred medium for reductions requiring solvated electrons;

S23

this is the electrochemical equivalent of a Birch reduction of the


benzene nucleus.76
A special medium exists in a cellular membrane where, between
the outer polar groups, there are non-polar aliphatic side chains.
Important redox systems, such as ubiquinone UQ, are embedded in
the membrane; phospholipids adsorbed on mercury have been used
to mimic a cellular membrane.77 Much debate has arisen on the
sensitivity of the electrochemical properties of UQ to the environment. The oxidation of UQH2 in a phospholipid model membrane
showed a bifurcation of the electron transfer necessary to achieve an
asymmetry across the membrane and to realize the required vectorial charge-transfer character.78
Organic solvents may be undesirable, especially for industrial
applications, and emulsions may present an alternative. Emulsions
may be simple emulsions water in oil or oil in water or a
bicontinuous microemulsion. 79,80 Such bicontinuous microemulsions are nontoxic, have a large interfacial area facilitating an intimate mixing of polar and nonpolar reagents, and high conductivity.
In aprotic media the choice of the anode material and anode
reaction is important, if an undivided cell is used; in aprotic solvents
a diaphragm is not without problems. The use of aluminum81 and
magnesium82 as sacrificial anodes solves not only the problem of
anode reaction, but the dissolved ions function as added supporting
electrolyte and in many cases the magnesium ions play a special role
by salt formation of the product during the reaction.
Electrode surfaces may be modified in many ways to get better
catalytic effects, chiral induction, or some other desirable effect not
obtainable at the classical electrode materials. Irreversible
adsorption,83 covalent attachment,84-86 and polymerization of monomers containing electroactive groups87,88 have been used. Voltammetric studies of the kinetics and energetics of coupled electrontransfer reactions in proteins adsorbed at pyrolytic graphite
electrodes have been made and it was found that the rates were more
likely to be controlled by the chemical reactions coupled to the
electron transfer than by the electron transfer.89 The modified electrodes are usually not stable enough for preparative reactions using
high current densities and are mostly employed as sensors;90 this
application will be discussed in another paper in this series.
Reductive coupling reactions.Many investigations on the mechanism of the hydrodimerization of activated olefins were inspired by the discovery by Baizer that on electrolysis under certain
conditions, acrylonitrile, an olefin activated by an electronwithdrawing group EWG, could dimerize to adiponitrile which
could be transformed to the two components, adipic acid and hexamethylenediamine, used for the synthesis of Nylon 66
2 CH2 CHCN 2 e 2 H NCCH2 CH2 CH2 CH2 CN
On electrolysis of an activated olefin a one-electron reduction to
a dimer or a two-electron reduction with saturation of the double
bond could take place, but several side-reactions, such as a Michael
addition of a nucleophile to the double bond or a polymerization
could compete with the electrochemical reactions. For the dimerization two main mechanistic possibilities presented themselves: a
dimerization of two radical species RR or an addition of a radical
species to the substrate RS followed by an electron transfer
RCHCH-EWG e
RCHCH-EWG
2 RCHCH-EWG
EWGCH-CHR-CHR-CH-EWG 2
followed by product-making steps. For certain activated olefins
e.g., some ketones a preprotonation takes place in protic solution,
so two neutral radicals dimerize. The radical-substrate dimerization
RS may be represented by

S24

Journal of The Electrochemical Society, 149 4 S21-S33 2002


RCHCH-EWG e
RCHCH-EWG
RCHCH-EWG RCHCH-EWG

EWGCH-CH-CH-CH-EWG
EWGCH-CH-CH-CH-EWG e/ RR CCR EWG

EWGCH-CH-CH-CH-EWG 2 RCHCH-EWG
followed by product-making steps.
The reactions are indicated to be reversible, but the degree of
reversibility depends on the competition between the backward reaction and the further reaction. The outcome of an electrolysis is a
function of the competition between different possible reactions and
thus on a number of rate constants. The progress in the elucidation
of such reactions was therefore dependent on the development of the
electroanalytical techniques and simulation capabilities. The dimerization of acrylonitrile was too fast for the electroanalytical techniques in the sixties and seventies, so other compounds with slower
kinetics were used for studying electrohydrodimerizations, not only
of activated olefins, but also other types of compounds, e.g., anthracenes substituted with EWG, which could dimerize reductively.
An early argument against R-R dimerization was that two radical
anions could not dimerize due to electrostatic repulsion, but using
different electroanalytical techniques and setting up working curves
for the expected theoretical behavior Bard,91 and Saveant92 could
show that the more slowly reacting 1,2-diactivated olefins in aprotic
medium dimerized reductively through a coupling between two radical anions.
However, the elucidation of a reaction mechanism for some activated olefins was made difficult by its dependence on the medium,
as this could influence the rate constants of the competing reactions
differently. An illustrative example is the hydrodimerization of an
alkyl cinnamate in superdry DMF, in DMF containing a small concentration of water, and in methanol. In superdry DMF the protonation of the dimerized dianion is not fast compared to the dissociation
of the dianion, whereas a small water concentration is enough to
ensure a fast protonation of the dianion and thus effectively an irreversible dimerization.93 In methanol a preprotonation may take
place and the dimerization occurs between two neutral radicals. If
the reversibility of the dimerization in the superdry medium is not
taken into account, the interpretation of the electroanalytical data
may lead to a false mechanism.
The discussion of the correct interpretation of experimental data
obtained by different research groups on the R-R vs. R-S problem
was in the beginning of the eighties rather heated. Since then further
development of instrumentation and power of simulation of theoretical curves have indicated that in general the hydrodimerization of
activated alkenes proceeds through a radical anion-radical anion
coupling.94,95
It was also early recognized that addition of a small concentration of water to the

aprotic medium increased the rate of the dimerization, but there was
discussion of the role of the water. In such media activated olefins,
such as derivatives of cinnamic acid esters, amides, nitriles, have
been found to hydrodimerize nearly exclusively to the d,l-form; the
esters made a ring closure to an all trans ring. The predominant

formation of the d,l-form has been explained by solvation,92


complexation,96 or hydrogen bonding.97,98 Hydrogen-bonding between a water molecule and the carbonyl groups of two radical
anions like a template was suggested to be essential for the predominant formation of the d,l-isomer of the dimer.97 In methanol and
other protic media approximately equal amounts of d,l- and mesoforms were formed.
Reduction of alkyl and benzyl halides in DMF to the anions in
the presence of activated olefins may lead to a Michael addition; the
product anion may react in different ways, such as protonation or
SN2 reaction with the starting halide.99 In the presence of carbon
dioxide carboxylic acids can be formed.100
The use of sacrificial anodes such as Al101 or Mg102 may be an
advantage in the carboxylation of both aliphatic and aromatic halides. Sacrificial anodes are also used in the carboxylation of terminal alkynes or diynes which on electrolysis at a silver cathode
formed dihydrogen and the substrate anion that reacted with the
carbon dioxide.103
Carboxylation may be catalyzed by nickel complexed with
Ph3 P104 or salen;105 electrolysis of unsaturated haloaryl ethers at a
carbon fiber cathode in DMF in the presence of NiII (cyclam) leads
to dihydrobenzofuran or dihydrobenzopyran derivatives in a tandem
cyclization-carboxylation.106

Another type of reductive coupling is intramolecular coupling;


the ring closure may result from a reaction between an electrochemically generated nucleophile and an electrophilic center or by a radical coupling.
Reduction in acetonitrile/Et4 NOTs with diethyl malonate as proton donor of an activated olefin having a carbonyl or cyano group in
an ortho position results in a ring closure by attack of the nucleophilic -carbon on the carbonyl group.107 A similar reductive ring
closure where the -carbon attacked a halide had been described
earlier.108 A slightly different type of ring closure takes place when
a carbonyl derivative with an unactivated double bond in the - or
-position is reduced with constant current in an undivided cell using about 10 F/mol. In this case, however, it is the carbonyl radical
anion or the oxygen-protonated radical anion which adds to the
double bond.109
The reductive formation of nucleophiles followed by an attack
on an electrophilic center is also the basis for the synthesis of many
heterocyclic compounds.110-112 Thus the reduction of the nitro group
in suitable ortho-substituted nitro compounds to the hydroxylamine
stage may lead to an N-oxide of an N-heterocyclic compopund;
thus, for example, o-nitrobenzalacetylacetone is reduced to
3-acetyl-2-methylquinoline-N-oxide.110

On reduction of aryl halides the initially formed radical anion


decomposes to halide ion and the aryl radical; the aryl radical is
generally more easily reduced than the substrate or it may abstract a
hydrogen atom from the solvent system, but in suitably substituted

Journal of The Electrochemical Society, 149 4 S21-S33 2002


compounds the radical may attack a phenyl ring with ring closure.
Thus, for example, 5-2-chlorophenyl-1-phenylpyrazole is reduced
to pyrazolo1,5-fphen-anthridine.113
Anodic substitution and coupling.Anodic reactions without added
mediator may involve a loss of electrons from the substrate or from
the solvent The previously mentioned pyridination of anthracene36
and the methoxylation of furan45 are examples of the first type; in
some oxidations the solvent is oxidized with loss of an electron and
a proton and the radical thus formed abstracts a hydrogen atom from
the substrate.114,115
Methoxylation of alkylbenzenes may take place in the side chain
or in the nucleus; 4-methylanisole can thus form methoxybenzyl
methyl ether or 1,1,4-trimethoxy-4-methylcyclohexa-2,5-diene depending on the medium; when the electrolysis is conducted under
conditions where formation of methoxy radicals also takes place,
nuclear methoxylation dominates.116 The methoxylation of methylbenzenes is of industrial importance; BASF oxidizes substituted
toluenes in methanolic solution at a carbon anode in a capillary-gap
cell to the substituted benzaldehyde dimethylacetal.

S25

product are formed, but by having a great excess of the cheaper


carboxylate good yields of the cross-coupled product may be obtained; such a reaction has been employed for the synthesis of
pheromones.124 The cross-coupled Kolbe reaction has been compared with low temperature photolysis of neat unsymmetrical diacyl
peroxides, and the latter was found to be better with regard to yield
and selectivity.125
Radicals produced in the Kolbe raction may also add to olefins,
and the radical thus formed can dimerize or be oxidized to a carbenium ion which can react with a nucleophile or may lose a proton to
an alkene. This type of addition of a radical to an alkene has been
used in ring-closure reactions126 and to introduce a trifluormethyl
group in different compounds.127,128

RC6 H4 CH3 4 e 2 CH3 OH RC6 H4 CH OCH3 2 4 H


Another useful reaction is the methoxylation of
N,N-dialkylamides, the first example being the oxidation of dimethylamides in MeOH/NH4 NO3 to N-alkoxy-N-methylamides; under these conditions an oxidation of both the amide and the nitrate
ion takes place.117 A large number of heterocyclic derivatives have
been prepared through the anodic methoxylation of acylated
heterocycles,118 in peptidomimetics,119 and in the first example of
combinatorial electrochemistry.120

Oxidation of aromatic compounds in acetic acid with some acetic


anhydride may result in acetoxylation; similar to the methoxylation,
attack on alkylarenes may occur both in the nucleus or in the side
chain. Substitution in the aromatic ring requires the presence of
acetate ion.120 When acetonitrile containing some trifluoracetic anhydride is the solvent, the carbocation formed by oxidation may
attack the solvent in a Ritter-type reaction; methylsubstituted benzenes are good substrates for such an acetamidation.121 Even
n-alkanes may be oxidized in MeCN/R4 NBF4 with acetamidation at
a secondary carbon.122

The further oxidation of the radicals to carbenium ions can lead


to acetoxylation, methoxylation or acetamidation, depending on the
solvent. The methoxylation reaction has been used in several syntheses, e.g., of carbohydrate derivatives129 or oligopeptides.130
Anodic oxidation of 2,4,6-tri-t-butylphenolate gives stable radicals, whereas 2,6-di-t-butylphenolates substituted in the 4-position
with a primary alkyl group dimerize reversibly to a para-substituted
quinol ether on anodic oxidation. In CV an irreversible peak is observed at fast sweep rates, whereas at slower sweep rates a cathodic
peak due to the reduction of the radical appears. This cathodic peak
becomes larger when the sweep rate is decreased due to the increased influence of the reversibility of the dimerization.131

Indirect electrolysis.In indirect electrolysis the electrons are not


exchanged heterogeneously directly between the electrode and the
substrate, but the electrode exchanges electrons with a compound, a

Ar-CH3 e
Ar-CH
3 H ArCH2 (e) ArCH2 MeCN ArCH2 -NC -Me H2 O ArCH2 NHCOMe H

The Kolbe reaction was the first useful electroorganic synthesis


and it is still under active investigation. In the classical Kolbe reaction the carboxylate is oxidized to a radical, but under certain conditions this radical may be oxidized further to a carbenium ion in the
so-called non-Kolbe reaction. The influence of solvent, current density, degree of neutralization and chain length on the selectivity of
the Kolbe vs. the non-Kolbe reaction has been clarified.123
RCOO e R CO2 1/2 R-R CO2
R e R Nu R-Nu
A useful extension of the classical reaction is the Kolbe crosscoupling reaction in which two carboxylates are oxidized simultaneously. Both the two symmetrical dimers and the cross-coupled

mediator, which then in the solution exchanges electrons with the


substrate. A theory of the dependence of the rate of electron transfer
on the driving force for outer-sphere electron transfer was developed
by Marcus132 for inorganic systems; later it was shown to be applicable also for organic compounds.133,134 The Marcus equation gives
the free energy of activation as a quadratic dependence on the driving force and the reorganization energy, the latter being the sum of
the solvent and structural reorganization energy necessary to reach
the transition state for the electron transfer. For relative small driving forces the rate of electron transfer increases with increasing
driving force, but when the driving force exceeds a certain value the
quadratic dependence in the Marcus equation leads to the counter
intuitive result that a further increase in driving force results in a
decrease in the rate of electron transfer. For several years attempts to

Journal of The Electrochemical Society, 149 4 S21-S33 2002

S26

experimentally verify this predicted result failed, but eventually it


was shown that experiments supported the theory.135,136 In the first
experiments the rate was measured of the electron transfer from a
biphenyl radical anion bound to one end of a steroid structure to an
aromatic compound A bound to the other end of the structure.

Indirect electrolysis can be a simple redox catalysis or it can


involve a chemical catalysis;137 in redox catalysis the mediator acts
merely as an electron donor or acceptor whereas an adduct between
the donor and acceptor which decomposes in a further reaction is
involved in chemical catalysis.
Inorganic mediators have been employed during the whole century, but the use of organic compounds as mediators began in the
middle of the seventies.138-141 Typically an aromatic compound is
reduced in an aprotic medium to the radical anion or dianion which
transfers an electron to a substrate. On electron uptake the substrate
forms a radical ion which undergo a fast, irreversible reaction, usually a cleavage, or it reacts by a dissociative electron transfer to a
radical and an anion. For an aromatic substrate the aryl radical is
usually reduced by the mediator to the anion, whereas the radical
from an aliphatic substrate has a possibility of being reduced by or
couple with the mediator. For an aromatic mediator, A, and an aliphatic halide, RX, the reactions can be described in a simplified way
as
A e
A
A RX A R X
kc

A R AR
kc

A R A R
The coupling reaction has been used for the synthesis of several
compounds,142,143 and the competition between coupling and reduction has been used, as described below, to estimate the potentials of
short-lived radicals.

Redox catalysis is in most cases based on the fact that the rate
constant for the heterogeneous electron transfer to nonconjugated
compounds is low due to the comparatively large changes in solvation energy and in bond-length and bond-angles associated with the
electron uptake and bond cleavage. An advantage of indirect electrolysis is that the reduction can take place at a potential less negative than that necessary for a direct reduction. The reduced mediator
diffuses into the bulk of the solution and the reaction with the substrate is then a homogeneous reduction; the indirect electrolysis can
thus be compared to a reduction taking place at a three-dimensional
electrode. The radicals which are formed in the solution by the electron transfer might add to alkenes or other unsaturated groups.143
Another advantage of redox catalysis is that very selective cleavage
and deprotection reactions may be performed. An example is the
cleavage of an allyl phenyl sulfone without cleavage of a homoallyl
phenyl sulfone in the same molecule144

A e
A
ArSO2 -CH R -C R CH-CH2 -CH2 -SO2 Ar 2 A H
ArSO
2 RCH2 C R CH-CH2 -CH2 -SO2 Ar 2 A
Macrocyclic nickel complexes have been used as catalysts in the
activation of organic halides. Electrogenerated NiI-complexes react rapidly with alkyl bromides and iodides, the reaction leads to
alkyl radicals and regeneration of the NiII-complex. In most cases
electrolysis of alkyl halides in the presence of catalytic quantities of
nickel complex leads to the complete reduction of the alkyl
halide.145
Recently nanometer-sized monolayer-protected clusters on gold
MPCs have been used in indirect electrolysis.146 MPCs are prepared by reducing a gold salt (AuCl
4 ) in a two-phase system in the
presence of an alkanethiol by NaBH4 ; gold particles of 1-3 nm
diameter bearing a surface coating of thiol are obtained. These particles are stable and can repeatedly be isolated by filtration from and
redissolved in organic solvents.147 MPCs may be prepared with
functional groups in the -position via a place-exchange reaction in
which the original alkanethiol is replaced by a thiol bearing the
required groups.148 Using anthraquinone bound to such monolayerprotected gold clusters, each bearing about 18 anthraquinone units,
indirect reduction of 1,1-dinitrocyclohexane was accomplished. The
catalytic efficiency of the clusters, the ratio between the current in
the presence of the substrate and the current of the mediator alone
divided by the concentration of the substrate, was found to be higher
than that of monomer anthraquinone which was attributed to the
smaller diffusion coefficient of cluster-bound anthraquinone relative
to that of monomer anthraquinone; this resulted in a spatial compression of the reaction zone next to the electrode. A review on
MPCs has been published.149
Self-assembled monolayers SAMs of alkylthiolates on gold
may also be used for immobilization of enzymes on an electrode;
this may avoid a partial denaturation of the enzyme often encountered with direct adsorption and still establish a short distance between the prosthetic group and the electrode surface. This approach
has been used in biosensors.150
Even living cells use the advantage of indirect electrolysis; extracellular electron transfer may be a general mechanism whereby
microorganisms generate energy for cell growth and/or
maintenance.151 Although mostly used for reducing metal salts, especially iron salts, reduction of organic compounds is also found, for
instance reduction of carbon tetrachloride.152
Indirect electrochemical oxidation using inorganic salts has been
used during the whole century, but more recently the use of organic
mediators has been employed. Anodic oxidation in acetonitrile or
dichloromethane of brominated derivatives of triphenylamine forms
a stable radical cation which may oxidize substrates such as thioethers. It thus selectively cleaves the methoxybenzylthio group in the
presence of other protecting groups such as trityl or tertbutyloxycarbonyl BOC.153 Other compounds such as polysubstituted biphenyls or dibenzothiophenes may also be used as mediators
in anodic indirect reactions.154

Some information on the electron exchange of redox enzymes


may be gained using CV of immobilized enzymes; such an immobilization should avoid adsorptive denaturation and improper orientation of the enzyme. One way is to use an antibody-antigen binding
procedure155 by which the full activity of the enzyme is preserved
and a long life-time achieved. The homogeneous enzymatic catalysis

Journal of The Electrochemical Society, 149 4 S21-S33 2002


of the oxidation of glucose to glucono--lactone by glucose oxidase
has been investigated by CV using ferrocenium methanol as mediator. The ferrocenium oxidizes FADH2 to FAD flavin adenine dinucleotide and FAD oxidizes glucose. From the catalytic increase
in CV of the immobilized enzyme in the absence and presence of
glucose information on the kinetics of the reactions can be
gained.156 A similar investigation with ferrocene attached at the end
of a long chain through an antibody to the supramolecular structure
mimicing a co-enzyme has been made.157
Radicals.In the second half of the 20th century radical chemistry
has been of steadily increasing importance. The redox potentials of
the radicals may help in predicting the course of such reactions. For
relatively long-lived radicals, conventional CV can measure the
thermodynamic potentials, but for short-lived radicals the rate of the
follow-up reaction and the rate of the heterogeneous electron transfer influence the measured potentials. At fast sweep-rates the
iR-drop becomes inconveniently large. For aliphatic radicals slow
electron transfer may shift the peak potential about 1 V from the
standard potential of the radical.158 A review of the measurement
and estimation of redox potentials of organic radicals has been
published.159
The introduction of ultramicroelectrodes54 in the nineteen eighties made it possible to outrun the follow-up reactions in many cases
and obtain the standard potentials as the midpoint between the cathodic and anodic signal. The distance between two potentials may
be larger than for a reversible reaction due to a slow electron transfer, but the rates of the forward and backward reaction are usually
comparable.
Radicals may be formed by photolysis and detected by voltammetry; a sensitive method of voltammetric detection was developed
by Wayner et al.160 The principle is that the intensity of the light,
and thus the concentration of the radicals formed, is modulated by a
mechanical chopper which blocks the light with a certain frequency.
Phase-sensitive detection of the voltammetric signals from the radicals filters off noise and other signals not in phase with the chopper.
By this means radicals may be detected in the 107 molar range
where second order follow-up reactions become less important. Often both the reduction and oxidation potentials of the radical can be
obtained
h

R-CO-R 2 R CO
h

t-Bu-O-O-t-Bu 2 t-BU-O
t-Bu-O RH t-Bu-OH R
Another way of measuring potentials of short-lived radicals is by
photoelectron injection which was developed by Russian
scientists.161 A metal electrode is illuminated with a short laser pulse
and low-energy electrons are injected into the solution and are solvated close to the electrode 10-60 . Most of the electrons diffuse
back to the electrode, but some may react with the substrate to form
radicals. Depending on the potential of the electrode the radicals
may or may not be reduced at the electrode, which influences the net
photoinduced charge. By doing a series of experiments at different
electrode potentials a kind of current-potential curve may be obtained from which the reduction potential of the radical can be
found. As the radicals have a very short distance to diffuse to the
electrode even potentials of short-lived radicals may be obtained.
Several types of radicals have been investigated by this
method.161,162
Indirect electrolysis of alkyl halides was discussed above and
these reactions were further developed by the Aarhus group to a
method for determination of potentials of short-lived radicals.163
The method is based on the competition between coupling of a
radical anion with a radical and reduction of the radical by the
radical anion. Aromatic compounds which in an aprotic medium

S27

form stable radical anions are chosen as electron donors. The radical
anion (A. ) transfers an electron to the substrate, typically an alkyl
halide RX, which then dissociates to the alkyl radical and halide
ion. The competition reactions are:
kc

A R AR
ke

A R A R
The competition is expressed by a competition parameter q
k e /(k e k c ); for q 0 the coupling dominates and for q close
to 1 the reduction of R . to R takes place. q-values for different
radical anions are measured and a plot of q vs. E Ao gives an S-shaped
curve. From the potential of the midpoint of the S-shaped curve,
q
E 1/2
, the standard potential, E Ro , can be calculated when the total
self-exchange reorganization energy R is known or estimated. k c is
assumed constant164,165 and about 109 M1 s1 . The weak point of
the method is the estimation of , but the method gives a reasonable
potential determination of short-lived radicals using easily available
apparatus.
Electrogenerated bases.When electrons are added to a molecule,
the resulting species may be characterized as a nucleophile, a base
and/or an electron donor depending on the follow-up reactions. In
most earlier reactions the anion/base generated during an electrolytic
reduction was protonated by the solvent system or an added proton
donor, so in protic solvents electrogenerated bases EGB were not a
problem. The first published reaction in which it was observed that
the EGB played a role in the reaction scheme was the reduction of
cyanoethylphosphonium toluenesulfonate in aprotic medium.
Baizer166 realized that the electron consumption n 1 observed in
the controlled potential reduction of the phosphonium salt in dry
DMF was caused by the deprotonation of the substrate by the anion
formed in the reduction, and that the anion of the substrate was not
reducible at the applied potential. Half of the substrate was thus not
reduced, but survived the reduction as the non-reducible anion. A
kinetic analysis of the self-protonation reaction a father-son reaction has been published by Saveant and Vianello.167
The first example of a synthetic application of an electrogenerated base using the type of compounds which later were called probases was provided by Lund.168 The strategy was to reduce a compound more easily reducible than the substrate components, so the
only function of the reduced probase was to abstract a proton from
one of the reaction partners. Azobenzene was reduced in a DMF
solution in which also benzaldehyde and benzyltriphenylphosphonium chloride were dissolved. In this solution azobenzene was the
most easily reduced species and the EGB formed the radical anion
of azobenzene and/or the monoanion of hydrazobenzene deprotonated the phosphonium salt, and the ylid thus formed reacted with
benzaldehyde in a Wittig reaction to form stilbene
PhNNPh e
PhNNPh
PhNNPh Ph3 P CH2 Ph
PhNH-N -Ph Ph3 P-CH-Ph
Ph P-CH-Ph PhCHO Ph3 PO PhCHCHPh
In DMF and DMSO rather strong EGBs can be made and these
bases may be so strong that they may replace conventional strong
bases such as BuLi, which have certain disadvantages for industrial
use. By using sterically hindered bases the nucleophilic reactions of
the EGB may be suppressed.169
An attractive way to produce an EGB is to reduce a compound
with a slightly acidic hydrogen at an electrode with a low hydrogen
overvoltage. This has been exploited by Shono170 who by direct
electrolysis of 2-pyrrolidone pKA (DMSO) 24.2 in DMF at a
platinum electrode formed the anion. As the hydrogen evolution

S28

Journal of The Electrochemical Society, 149 4 S21-S33 2002

requires a rather negative cathode potential, the base had in most


cases to be formed before the addition of the substrate.
Carbon bases may be formed by reduction of a halogen compound. An example of this is the reduction of a low concentration of
carbon tetrachloride in the presence of benzaldehyde and a high
concentration of chloroform. The carbanion formed on reduction of
carbon tetrachloride adds to benzaldehyde and the adduct deprotonates chloroform to the reactive carbanion171

CCl4 2 e CCl
3 Cl

RCHO CCl
3 RCH O CCl3

RCH O CCl3 HCCl3 RCH OH CCl3 CCl


3
Dioxygen is reduced to superoxide anion which is a relatively
weak base, but on protonation two superoxide anions can form the
172
stronger base HO
Dioxygen may be used as EGB,173 but the
2 .
substrate anion is likely to react with dioxygen to a peroxide. This
has been avoided by using a low concentration of oxygen in a deprotonation of nitromethane which was used in a condensation of nitromethane with aldehydes. By having nitromethane as solvent, the
competition between deprotonation and reaction with dioxygen favored the deprotonation.174 An alternative reaction of superoxide
anion has been suggested in the reduction of oxygen in the presence
of 4-nitrotoluene; the superoxide anion abstracts a hydrogen atom
from the methyl group and the benzylic radical is oxidized by
dioxygen.175
Aromatic nucleophilic substitution.The radical chain nucleophilic
substitution, SRN1, discovered by Kornblum, Russell and Bunnett,
has been investigated by Saveant et al.176 by electrochemical techniques. The reaction involves the following steps where AX is an
aromatic compound substituted with a leaving group X and Nu is a
soft i.e., easily oxidizable nucleophile
ArX e
ArX
ArX
Ar. X
Ar Nu
ArNu
ArNu ArX
ArNu ArX
The cathodic reduction of the radical or a hydrogen abstraction
from the solvent are terminating reactions. Liquid ammonia has
mostly been used as a solvent to minimize hydrogen abstraction
from the solvent by the aryl radical.
Aromatic radical anions (ArX. ) have a certain lifetime that
gives them time to diffuse from the electrode and thus avoid reduction of the relatively easily reducible radical. In some cases an indirect reduction of ArX may be an advantage.177,178 The cleavage of
the substrate radical anion to an aryl radical and a nucleophile X is
the opposite reaction of the coupling of the nucleophile, Nu, with
the aryl radical. Differences in nucleophilicity and concentration determine the course of the reaction.
Aliphatic halides undergo as a rule not an electrochemically induced SRN1 reaction, as the radical is formed close to the electrode
in a dissociative electron transfer and is thus reduced to the anion
before it can react with the nucleophile. An interesting case of a
benzylic halide is the classical reaction between 4-nitrocumyl chloride and 2-nitropropanate ion which reacts in a chain reaction without external stimulation in a thermal SRN1 reaction. Saveant179
showed that an outer-sphere electron transfer from the anion to the
substrate was thermodynamically unfavored, but that a decrease in
driving force was able to change the mechanism of the reductive
cleavage of the substrate from a stepwise to a concerted reaction.
Aliphatic nucleophilic substitution.Indirect electrolysis has been
used to investigate the mechanism of the aliphatic nucleophilic substitution which traditionally has been regarded as following the clas-

sical SN1 and SN2 reactions or hybrids between them. In the investigation of a possible electron transfer mechanism in the
substitution, the Aarhus group used a nucleophile, the anion of 1,4dihydro-4-methoxycarbonyl-1-methylpyridine, which on electrochemical oxidation gave a stable radical, so the reversible oxidation
potential of the nucleophile could be found by CV. Other similar
nucleophiles were also used. The method consisted in a comparison
of the rate of reaction of the nucleophile with an alkyl halide with
the rate of electron transfer from a stable radical anion with the same
oxidation potential as the nucleophile to the same alkyl halide. For
sterically hindered alkyl halides the two rates were practically the
same, indicating that in this aliphatic nucleophilic substitution the
rate of the transfer of a single electron from the nucleophile to the
alkyl halide was the rate-determining step.180,181 In less sterically
hindered alkyl halides the rate of substitution was faster than the
expected rate of electron transfer indicating either a competition
between electron transfer and SN2 or a kind of hybrid between the
two mechanisms.

Chiral induction.Several different ways have been explored for


the transfer of chirality during electrochemical reactions. Horner182
used chiral supporting electrolytes, but obtained rather modest chiral
induction. Seebach183 used a chiral solvent, S,S()2,3dimethoxy-1,4-bis(dimethylamino)butane, with no better result.
Grimshaw184 employed alkaloids adsorbed at the electrode as chiral
proton donors, but the enantiomeric excess (ee) was not above 20
%. Other groups have attempted to improve the induction185-188 using adsorbed alkaloids and fairly high induction has been obtained
in some cases.
Chemically modified electrodes have been used to introduce
chirality. Miller189 modified the cathode with (S)-phenylglycine and
obtained a modest ee in the reduction of prochiral ketones.
Nonaka190 prepared different electrodes coated with optically active
polyaminoacids and got variable and sometimes very high ee both
in reductions and oxidations, and Osa191 used an anode modified
with TEMPO in the presence of ()-sparteine in an oxidative coupling.
An attractive way to obtain stereoselective reactions would be to
use enzymatic reactions in which the active form of the enzyme was
regenerated electrochemically. Enzymes, like most proteins, exchange electrons slowly with an electrode, so to get a useful synthetic route this difficulty must be overcome. Another problem
might be to obtain long-time stability of the enzyme necessary for
preparative applications. Different approaches to solve these problems have been published.
A redox-enzyme usually requires a co-factor and in most cases a
mediator is necessary to ensure a fast exchange of electrons between
the electrode and the co-factor. One approach has been to immobilize the enzyme, its co-factor, and a mediator at a polyacrylic acidcoated graphite-felt electrode. The effective surface area is rather
small and the reactions are not fast so the productivity is limited by
slow reaction steps and not by mass transport.192
A way to increase the effective surface area is to deposit a fibrous
layer of polyaniline on reticulated vitreous carbon and adsorb the
enzyme on the polyaniline. The enzyme was stabilized by crosslinking with glutaraldehyde and then immobilized with a thin layer
of poly1,2-diaminobenzene. By this increase in the effective surface layer the current density could, for a not too high substrate
concentration, have some dependence on the mass transport rate,
and the cross-linking of the enzyme with glutaraldehyde increased
considerably the stability of the enzyme.193

Journal of The Electrochemical Society, 149 4 S21-S33 2002


A further possibility is to use an ultrafiltration membrane with a
molecular weight cutoff so that the enzyme cannot penetrate it. This
approach requires that a redox catalyst and the co-factor be bound to
a soluble polymer such as polyethylene glycol. This procedure has
been called homogeneous immobilization by Steckhan,194,195 and
it avoids the diffusional limitations at the electrodes with immobilized enzymes. In the enzyme membrane reactor the product diffuses
through the membrane, whereas the enzyme, the mediator and the
co-factor, all bound to polyethylene glycol which cannot diffuse
through the membrane, are recirculated. The reaction is controlled
by proper adjustment of the residence time in the reactor. This approach might be the best route for development of electroenzymatic
processes for industrial use.
Polymers.Electrogenerated nucleophiles/anions have been shown
to initiate polymerization of suitable monomers; in the absence of
impurities each electron transfer starts a chain. Control of chain
length may be done by addition of a small amount of proton donor
or by interposing a pattern of negative and positive pulses separated
by a chosen time interval. Such a pattern of pulses initiates a chain,
lets it grow to a certain length, and the chain is terminated by the
positive pulse. Examples of monomers electrochemically polymerized are styrene,196 phenylacetylenes,197 propylene oxide,198 and
methyl methacrylate.199
The finding that polyacetylene on oxidation with iodine changed
the polymer from an insulator to a material with nearly metallic
conductivity200 initiated much research on conducting polymers, at
first primarily to develop a rechargeable battery. It was found that
besides polyacetylene,201 polypyrrole,202 polythiophene,203 and
polyaniline204 could be made electrical conducting by doping
which in this context is a redox reaction; p-doping corresponds to an
oxidation and n-doping to a reduction.
Conducting polymers differ from redox polymers in that in conducting polymers the monomers interact via a -electron system
whereas the chain in a redox polymer is not conjugated.
The polymerization of pyrrole begins with an oxidation to the
radical cation; two of these dimerize with a rate constant205 of about
109 M1 s1 and the dimerized species loses two protons. The
dimerization takes place at the -positions of the radical cations.
The radical cation of the dimer and especially the tetramer reacts
slower, so a radical cation reacts faster with another radical cation
than with the dimer or tetramer. The rate of the dimerization206 of
the radical cation of thiophene tetramer is only about 105 M1 s1 .
These results and results obtained by Heinze207 with
3-methylthiophene indicate that the

polymerization does not take place by chain propagation, adding a


monomer to a growing chain.
The charging process of conducting polymers has been suggested
to involve two coexisting subsystems at the polymer matrix; one
system to exchange electrons with the electrode in a quasi-reversible
manner and another system which involves chemically reversible
intermolecular bonds between neighboring polymer molecules.208
On oxidation or reduction of the insulating polymer to a conducting polymer, the polymer get a charge which requires the diffusion
of counter ions. This has to be taken into account in practical applications and in all theoretical treatments of the electropolymerization.

S29

Besides applied research for the development of batteries, research on conducting polymers for electrochromic displays ECD
has been continuing. An ECD is based on an electrochemical reaction of a material that displays a visual change on changing its redox
state, such as for instance, polypyrrole209 and polyaniline.210 By
using polyaniline with poly-o-phenylenediamine and polymetanilic
acid, the three primary colors could be developed.211,212
Another type of conducting polymer is DNA; electron vacancies
holes formed by oxidants such as hydroxyl radicals or singlet oxygen may, in spite of the catalase that is present, attack DNA; such
oxidative damage may contribute to ageing of cells, mutagenesis,
and cancer. Holes may, however, be transferred across up to 50 base
pairs in a DNA duplex in homogeneous solution by hopping between cytosine-guanine CG base pairs213 to a remote G which is
oxidized.214-217 When not aligned in parallel, condensed-phase DNA
duplexes are electrically insulating, but when aligned their onedimensional conductivity in the direction of the long axis is only
two orders of magnitude lower than that of iron. The increase in
conductivity is attributed to the high unidirectional polarizability of
DNA.218
It has been suggested that the high ability to conduct holes is
developed to protect essential parts of DNA against oxidation by
transferring the attacking hole to a less essential part of the genome.
Of the four bases guanine is more easily oxidized than the three
others E pH 7: G 1.04 V, T 1.29 V, A 1.32 V, and C 1.44 V vs.
NHE219,220 and by having long stretches of GC without essential
function in the cell and good conductivity, the damaging hole may
be transferred to the more easily oxidized part of the genome, and
the hole may end by oxidizing guanine to 7,8-dihydro-8oxoguanine. G is not only the most easily oxidizable base, but its
catalytic one-electron oxidation in poly-G sequences is particular
fast. The sacrificially oxidized protective domain may subsequently
be excised and repaired.221 The function of the sacrificial stretches
of guanine to protect the essential part of the genome has been
compared to the cathodic protection of iron by zinc; a review of this
hypothesis has recently appeared.222
Metalorganic compounds.Organoelemental compounds exist of
many elements of the periodic table and only a few investigations
can be presented here. Reviews of the electrochemistry of such compounds have been published.223-225
One of the most investigated organometallic compounds is ferrocene; it is reversibly oxidized to ferrocenium, and as the solvation
energy of the system is low, it has been recommended as a standard
system to compare potentials in different solvent systems.226 Ferrocene has been prepared electrochemically using an iron anode and
cathodically generated cyclopentadiene anion.227 Metallocenes may
be used in indirect reactions, for instance as a mediator for the
regeneration of FAD.
Metal complexes of porphine-type compounds, the so-called pigments of life, have been investigated intensively; the electrochemistry of metalloporphyrins in nonaqueous media has been reviewed by
Kadish.228
The cobalt-complex vitamin B12 and related compounds have
been investigated by Saveant229 by different techniques with determination of rate constants for cleavage and formation reactions and
constants for complex formation. Vitamin B12 has preparatively
been used for the catalysis of different reactions by Scheffold.230
Reduction of the CoIII-derivative of vitamin B12 cleaves the axial
ligand with further reduction to the CoI-derivative. This compound
can add an alkyl halide RX in an oxidative addition. The RCoIII
compound loses on reduction the radical R which may add to an
activated alkene; oxidative addition of RX to the CoI derivative
forming RCoIII completes the catalytic circle. The product is thus
akin to a Michael addition of R to the alkene.
A similar catalytic circle in which primary and secondary alkyl
radicals, formed by reducing a RNiII complex, are added to an
activated olefin may be achieved; oxidative addition of RX to the
Ni0 complex reforms the RNiII complex and completes the cata-

S30

Journal of The Electrochemical Society, 149 4 S21-S33 2002

lytic circle.231 Another reaction with carbon-carbon bond formation


is the Pd(PPh3 ) 4 -catalyzed synthesis of biaryls from aryl bromides
and iodides.232,233 and the CuII salt-catalyzed reaction of alkenylstananes to symmetrical dienes in a homocoupling with high
stereospecificity.234
Silicon and tin compounds have been used in different contexts.
Silicon interacts effectively with a neighboring -orbital or p-orbital
at a heteroatom and increases the HOMO level and thus lowers the
oxidation potential. Oxidation of an allylsilane proceeds smoothly
with cleavage of the C-Si bond and further oxidation of the radical
to the cation which reacts with a nucleophile in both allylic
positions.235 The silanes may be prepared by reduction of an organic
halide in the presence of chlorosilane236,237
RCHC R -CH2 SiMe3 e RCHC R CH
2 Nu
RCH Nu CR CH2
RCHC R CH2 Nu
When a stannyl and a silyl group are situated at the same carbon
atom, the CuSn bond is cleaved at a lower oxidation potential than
the CuSi bond. The CuSi bond may then be cleaved by raising
the potential and another nucleophile can be introduced.238 When a
Si or Sn group has been used to modify the potential, it has been
termed an electroauxilary group.
Industrial processes.In the mid-sixties two commercial processes
were introduced: the Nalco process for the production of tetraalkyl
lead239 and Monsantos production of adiponitrile240 by reductive
dimerization of acrylonitrile, the latter process developed by Manuel
M. Baizer. The successful introduction of these two productions
increased substantially the interest of chemical companies to consider development of industrial electrochemical reactions.
The key to the success of the Monsanto process was Baizers
discovery that the use of tetraalkylammonium salts as supporting
electrolyte in water as solvent created a layer at the cathode in which
the proton activity was low enough to avoid a protonation of the
radical anion faster than its dimerization but high enough to protonate the dimerized product and thus avoid a base-catalyzed polymerization of acrylonitrile.
The first Monsanto process used a divided cell with lead electrodes the anode with 1 % Ag and an ion-exchange membrane as
separator, but it was soon realized that it was possible to use an
undivided cell which improved the economy, although the selectivity was slightly lower 88 vs. 92 and the current density had to be
halved. However, the electrode distance was diminished from 7 to
1.8 mm, the necessary voltage was lowered from 11.6 V to 3.8 V, so
the energy consumption declined from 6700 to 2500 kWh/ton. The
cathode material was cadmium with carbon steel as anode. The electroreductive dimerization of acrylonitrile is now carried out by several companies and the volume of the production is about 300 000
t/yr.
In the Nalco process a Grignard reagent is made from an alkyl
chloride and magnesium in an ether solution, and the Grignard reagent, RMgCl, is oxidized at a sacrificial lead anode. The radicals
attack the lead with formation of tetraalkyllead. At the cathode magnesium salt is reduced to magnesium which reacts with the alkyl
chloride to form the Grignard reagent. The process was run for
about 20 years, but was abandoned in the early nineteen eighties due
to the diminished use of tetraalkyllead in gasoline
4 Mg 4 EtCl 4 EtMgCl
4 EtMgCl 4 e 4 Cl 4 Et 4 MgCl2
4 Et. Pb Et4 Pb
MgCl2 2 e Mg 2 Cl

The use of divided cells with membranes causes many problems


when non-aqueous solvents have to be employed, and sacrificial
anodes in an undivided cell may be an alternative. This has been
used in the production of Fenoprofen which is prepared by reduction
of chloroethyldiphenyl ether in N,N-dimethylformamide under a
carbon dioxide pressure of 2 bar. The carbanion formed in the reduction reacts with carbon dioxide to form the carboxylic acid.241,242
C6 H5 -O-C6 H4 -CHCl-CH3 2 e CO2
C6 H5 -O-C6 H4 -CH COO -CH3 Cl
The dream of an organic electrochemist has always been to make
useful products simultaneously at the cathode and the anode. Some
laboratory examples have been published,243 but only recently has
an industrial paired synthesis been realized by BASF.244 In an
undivided cell using methanol as solvent, phthalic acid dimethyl
ester is cathodically reduced to phthalide and at the anode 4-tertbutyltoluene is oxidized to 4-tert-butylbenzaldehyde dimethylacetal;
the methanol released at the cathode is used at the anode to form the
dimethylacetal. Previously phthalic acid anhydride was reduced
catalytically to phthalide and the anodic oxidation of 4-tertbutyltoluene gave dihydrogen as biproduct; the paired synthesis
avoids high pressure catalytic hydrogenation and the energy demand
has been drastically reduced.

References
1.
2.
3.
4.
5.
6.
7.
8.
9.
10.
11.
12.
13.
14.
15.
16.
17.
18.
19.
20.
21.
22.
23.
24.
25.
26.
27.
28.
29.

A. G. A. Volta, J. Nat. Phil. Chem. Arts, 4, 179 1800.


M. Faraday, Ann. Phys. (Leipzig), 47, 438 1834.
J. Kolbe, J. Prakt. Chem., 41, 138 1847.
A. Crum Brown and J. Walker, Liebigs Ann. Chem., 261, 107 1891; A. Crum
Brown and J. Walker, Liebigs Ann. Chem., 274, 41 1893.
Ch. F. Schoenbein, Liebigs Ann. Chem., 54, 164, 176 1845.
F. Haber, Z. Elektrochem. Angew. Phys. Chem., 5, 235 1898.
F. Fichter, Organische Elektrochemie, Steinkopff, Dresden 1942.
S. Swann, Jr., Trans. Electrochem. Soc., 69, 287 1936; S. Swann, Jr., Trans.
Electrochem. Soc., 77, 459 1940.
M. Le Blanc, Z. Elektrochem. Angew. Phys. Chem., 7, 292 1900.
R. L. Taylor, Chem. Metall. Eng., 44, 588 1937.
J. Heyrovsky, Chem Listy, 16, 256 1922; J. Heyrovsky, Philos. Mag., 45, 303
1923.
M. Shikata, Trans. Faraday Soc., 21, 42, 53 1925.
J. Heyrovsky and M. Shikata, Recl. Trav. Chim. Pays-Bas., 44, 496 1925.
J. Heyrovsky and I. Smoler, Collect. Czech. Chem. Commun., 4, 521 1932.
A. Winkel and G. Proske, Ber. Dtsch. Chem. Ges. B, 69, 693, 1917 1936.
G. Semerano and A. Chisine, Gazz. Chim. Ital., 66, 504 1936.
R. Brdicka, Collect. Czech. Chem. Commun., 5, 238 1933; R. Brdicka, Collect.
Czech. Chem. Commun., 8, 366 1936.
I. M. Kolthoff and S. C. Miller, J. Am. Chem. Soc., 62, 2171 1940.
I. M. Kolthoff and J. J. Lingane, Polarography, Interscience, New York 1941.
J. Heyrovsky, Polarographie, Springer, Wien, 1941.
H. A. Laitinen and S. Wawzonek, J. Am. Chem. Soc., 64, 1765 1942; S. Wawzonek and H. A. Laitinen, J. Am. Chem. Soc., 64, 2365 1942.
M. von Stackelberg and W. Stracke, Z. Elektrochem. Angew. Phys. Chem., 53, 118
1949.
F. Petru, Collect. Czech. Chem. Commun., 12, 620 1947.
R. M. Elofson, R. L. Edsberg, and P. A. Mecherly, J. Electrochem. Soc., 97, 166
1950
P. C. Tompkins and C. L. A. Schmidt, J. Biol. Chem., 143, 643 1942; P. C.
Tompkins and C. L. A. Schmidt, Univ. Calif. Pub. Physiol., 8, 237, 247 1944.
J. T. Stock, J. Chem. Soc., 1944, 427; J. T. Stock, J. Chem. Soc., 1949, 763.
J. J. Lingane, C. G. Swain, and M. Fields, J. Am. Chem. Soc., 65, 1348 1943.
A. Hickling, Trans. Faraday Soc., 38, 27 1942.
J. J. Lingane and S. L. Jones, Anal. Chem., 22, 1169 1950.

Journal of The Electrochemical Society, 149 4 S21-S33 2002


30. J. Heyrovsky and J Kuta, Grundlagen der Polarographie, Akademie-Verlag, Berlin 1965.
31. I. M. Kolthoff and J. J. Lingane, Polarography, 2nd ed., Wiley-Interscience, New
York 1952.
32. L. Meites, Polarographic Techniques, 2nd ed., Wiley-Interscience, New York
1965.
33. M. Brezina and P. Zuman, Polarography in Medicine, Biochemistry and Pharmacy, Wiley-Interscience, New York 1958.
34. P. Zuman, Substituent Effects in Organic Polarography, Plenum, New York
1967.
35. H. Matsuda and Y. Ayabe, Z. Elektrochem., 59, 494 1955.
36. H. Lund, Acta Chem. Scand., 11, 491, 1323 1957.
37. J. H. Simons, J. Electrochem. Soc., 95, 47 1949.
38. H. Meinert, R. Fackler, J. Mader, P. Reuter, and W. Roehlke, J. Fluorine Chem.,
51, 53 1991.
39. S. Rudiger, A. Dimitrov, and K. Hottmann, J. Fluorine Chem., 76, 155 1996.
40. P. Sartori and N. Ignatev, J. Fluorine Chem., 87, 157 1998.
41. V. Childs, C. Gross, and S. Polson, Abstract 1006, The Electrochemical Society
Meeting Abstracts, Seattle, WA, May 2-6, 1999.
42. I. N. Rozhkov and I. L. Knunyants, Dokl. Akad. Nauk SSSR, 199, 614 1971; I.
N. Rozhkov and I. L. Knunyants, Chem. Abstr., 76, 7291 1972.
43. H. Schmidt and H. D. Schmidt, Chem. Tech. (Leipzig), 5, 454 1953; H. Schmidt
and H. D. Schmidt, J. Prakt. Chem., 2, 105 1955.
44. T. Fuchigami, in Organic Electrochemistry, 4th ed., H. Lund and O. Hammerich,
Editors, p. 1035, Marcel Dekker, New York 2001.
45. N. Clauson-Kaas, F. Limborg, and K. Glens, Acta Chem. Scand., 6, 531 1952;
N. Clauson-Kaas and Z. Tyle, Acta Chem. Scand., 6, 962 1952.
46. S. W. Feldberg in Electroanalytical Chemistry, Vol. 3, A. J. Bard, Editor, p. 199
1969; M. Rudolph, D. P. Reddy, and S. W. Feldberg, Anal. Chem., 66, A589
1994.
47. D. Britz, Digital Simulation in Electrochemistry, 2nd ed., Springer, Berlin 1988.
48. A. J. Bard and L. R. Faulkner, Electrochemical Methods, John Wiley & Sons,
New York 1999.
49. C. P. Andrieux and J.-M. Saveant, in Investigations of Rates and Mechanisms of
Reaction, C. F. Bernasconi, Editor, p. 305, John Wiley & Sons, New York 1986.
50. S. U. Pedersen and K. Daasbjerg, Electrochemical Techniques in Electron Transfer in Chemistry, Vol. 1, Z. B. Balzani, Editor, John Wiley & Sons, New York
2001.
51. O. Hammerich, in Organic Electrochemistry, 4th ed., H. Lund and O. Hammerich,
Editors, Marcel Dekker, New York 2001.
52. J.-M. Saveant and E. Vianello, C. R. Acad. Sci. URSS, 256, 2597 1963.
53. R. S. Nicholson and I. Shain, Anal. Chem., 36, 706 1964; R. S. Nicholson and I.
Shain, Anal. Chem., 37, 178, 190 1965; R. S. Nicholson, Anal. Chem., 37, 1351
1965.
54. J. C. Imbeaux and J.-M. Saveant, J. Electroanal. Chem., 28, 325 1970; D.
Garreau and J.-M. Saveant, J. Electroanal. Chem., 86, 63 1978.
55. S. Antonello, M. Musumeci, D. D. M. Wayner, and F. Maran, J. Am. Chem. Soc.,
119, 9441 1997.
56. M. Fleishman, S. Pons, R. Rolison, and P. P. Schmidt, Ultramicroelectrodes,
Datatech Systems, New York 1987.
57. C. P. Andrieux, D. Garreau, P. Hapiot, and J.-M. Saveant, J. Electroanal. Chem.,
248, 447 1988.
58. D. O. Wipf and R. M. Wightman, J. Phys. Chem., 93, 4286 1989.
59. C. Amatore, E. Maisonhaute, and G. Siminneau, Electrochem. Commun., 2, 81
2000; C. Amatore, E. Maisonhaute, and G. Siminneau, J. Electroanal. Chem.,
486, 141 2000.
60. F. Zhou and A. J. Bard, J. Am. Chem. Soc., 116, 393 1994.
61. D. A. Treichel, M. V. Mirkin, and A. J. Bard, J. Phys. Chem., 98, 5751 1994.
62. F.-R. Fan and A. J. Bard, Science, 267, 871 1995.
63. T. Solomon and A. J. Bard, J. Phys. Chem., 99, 17487 1995.
64. A. L. Parker, P. R. Unwin, and J. Zhang, Electrochem. Commun., 3, 372 2001.
65. B. Liu, S. A. Rotenberg, and M. V. Mirkin, Proc. Natl. Acad. Sci. U.S.A., 97, 9855
2000.
66. M. V. Mirkin, B. Liu, and S. A. Rosenberg, 22nd Sandbjerg Meeting Abstracts, p.
30 2001.
67. C. Amatore, S. Arbault, D. Bruce, P. de Oliveira, M. Erard, and M. Vuillaume,
Faraday Discuss., 116, 319 2000.
68. L. Eberson, M. P. Hartshorn, J. J. McCullough, O. Persson, and F. Radner, Acta
Chem. Scand., 52, 1024 1998; L. Eberson, M. P. Hartshorn, F. Radner, and O.
Persson, J. Chem. Soc., Perkin Trans. 2, 1998, 59.
69. S. Wawzonek and M. E. Runner, J. Electrochem. Soc., 99, 457 1952; I. M.
Kolthoff and J. F. Coetzee, J. Am. Chem. Soc., 79, 870, 1852 1958.
70. J. Phelps, K. S. V. Santhanam, and A. J. Bard, J. Am. Chem. Soc., 89, 1752 1967.
71. O. Hammerich and V. D. Parker, Electrochim Acta., 18, 537 1973.
72. E. Garcia and A. J. Bard, J. Electrochem. Soc., 137, 2752 1990.
73. R. S. Glass and V. V. Jouikov, Tetrahedron Lett., 40, 6357 1999.
74. N. S. Moe, Acta Chem. Scand., 21, 1389 1967.
75. C. Amatore, M. A. Oturan, J. Pinson, J.-M. Saveant, and A. Thiebault, J. Am.
Chem. Soc., 107, 3451 1985; C. Combellas, F. Kanoufi, and A. Thiebault, J.
Electroanal. Chem., 407, 195 1996.
76. R. A. Benkeser and E. M. Kaiser, J. Am. Chem. Soc., 85, 2858 1963.
77. C. A. Nelson and A. Benton, J. Electroanal. Chem., 202, 253 1986.
78. G. J. Gordillo and D. J. Schiffrin, J. Chem. Soc., Faraday Trans., 90, 1913 1994;
G. J. Gordillo and D. J. Schiffrin, Faraday Discuss., 116, 89 2000.
79. G. N. Kamau and J. F. Rusling, Langmuir, 12, 2645 1996; X. Zu and J. F.
Rusling, Langmuir, 13, 3693 1997.

S31

80. H. Carrero, J. Gao, J. F. Rusling, C.-W. Lee, and A. J. Fry, Electrochim. Acta, 45,
503 1999.
81. G. Silvestri, S. Gambino, and G. Filardo, Adv. Polym. Sci., 38, 28 1981; G.
Silvestri, S. Gambino, and G. Filardo, Acta Chem. Scand., 45, 987 1991.
82. O. Sock, M. Troupel, and J. Perichon, Tetrahedron Lett., 26, 1509 1985; C.
Saboureaum, M. Troupel, and J. Perichon, J. Appl. Electrochem., 20, 97 1990.
83. R. F. Lane and A. T. Hubbard, J. Phys. Chem., 77, 1401, 1411 1973.
84. B. F. Watkins, J. R. Behling, E. Kariv, and L. L. Miller, J. Am. Chem. Soc., 97,
3549 1975.
85. J. C. Lennox and R. W. Murray, J. Electroanal. Chem., 78, 395 1977; J. R.
Lenhard and R. W. Murray, J. Am. Chem. Soc., 100, 7870 1978.
86. C. A. Koval and F. C. Anson, Anal. Chem., 50, 223 1978.
87. G. Bidan, A. Deronzier, and J.-C. Moutet, J. Chem. Soc. Chem. Commun., 1984,
1185; S. Cosnier, A. Deronzier, and J.-C. Moutet, J. Electroanal. Chem., 193, 193
1985.
88. J. G. Eaves and H. S. Munro, J. Chem. Soc. Chem. Commun., 1985, 684.
89. F. A. Armstrong, R. Camba, H. A. Heering, J. Hirst, L. J. C. Jeuken, A. K. Jones,
C- Leger, and J. P. McEvoy, Faraday Discuss., 116, 191 2000.
90. E. A. G. Hall, Biosensors, Prentice-Hall, Upper Saddle River, NJ 1991.
91. V. J. Puglisi and A. J. Bard, J. Electrochem. Soc., 119, 829 1972; V. J. Puglisi
and A. J. Bard, J. Electrochem. Soc., 120, 748 1973.
92. E. Lamy, L. Nadjo, and J.-M. Saveant, J. Electroanal. Chem., 42, 189 1973; E.
Lamy, L. Nadjo, and J.-M. Saveant, J. Electroanal. Chem., 50, 141 1974.
93. R. D. Grypa and J. T. Maloy, J. Electrochem. Soc., 122, 509 1975; B. M. Bezilla
and J. T. Maloy, J. Electrochem. Soc., 126, 579 1979.
94. J.-M. Saveant, Acta Chem. Scand., B37, 365 1983; C. Amatore, J. Pinson, and
J.-M. Saveant, J. Electroanal. Chem., 137, 143 1982.
95. V. D. Parker, Acta Chem. Scand., B35, 147,149 1981; V. D. Parker, Acta Chem.
Scand., B37, 393 1983; O. Hammerich and V. D. Parker, Acta Chem. Scand.,
B37, 379 1983.
96. V. D. Parker, Acta Chem. Scand., B35, 147 1981.
97. I. Fussing, O. Hammerich, A. Hussain, M. F. Nielsen, and J. H. P. Utley, J. Chem.
Soc., Perkin Trans. 2, 1996, 649.
98. I. Fussing, O. Hammerich, A. Hussain, M. F. Nielsen, and J. H. P. Utley, Acta
Chem. Scand., 52, 328 1998.
99. M. M. Baizer and J. L. Cruma, J. Org. Chem., 37, 1951 1972.
100. J. H. Wagenknecht, J. Electroanal. Chem., 52, 489 1974.
101. G. Silvestri, S. Gambino, G. Filardo, and A. Gulotta, Angew. Chem. Int. Ed. Engl.,
23, 979 1984.
102. J. F. Fauvarque, Y. de Zelicourt, C. Amatore, and A. Jutand, J. Appl. Electrochem.,
20, 338 1990.
103. F. Koster, E. Dinjus, and E. Dunach, Eur. J. Org. Chem., 2001, 2507.
104. M. Troupel, Y. Rollin, J. Perichon, J. Fauvarque, Nuov. J. Chim., 5, 621 1981.
105. A. Gennaro, A. A. Isse, and F. Maran, J. Electroanal. Chem., 507, 124 2001.
106. S. Olivero and E. Dunach, Eur. J. Org. Chem., 1999, 1885.
107. D. P. Fox, R. D. Little, and M. M. Baizer, J. Org. Chem., 50, 2202 1985.
108. N. T. Nugent, M. M. Baizer, and R. D. Little, Tetrahedron Lett., 23, 1339 1982.
109. T. Shono and M. Mitani, J. Am. Chem. Soc., 93, 5284 1971.
110. H. Lund and L. G. Feoktistov, Acta Chem. Scand., 23, 3482 1969.
111. H. Lund, in Advances in Heterocyclic Chemistry, A. R. Katritzky and A. J. Boulton, Editors, p. 12, Academic Publishers, New York 1970; H. Lund and I. Tabakovic, in Advances in Heterocyclic Chemistry, A. R. Katritzky and A. J. Boulton,
Editors, p. 36, Academic Publishers, New York 1970; H. Lund in Organic Electrochemistry, H. Lund and O. Hammerich, Editors, p. 669, Marcel Dekker, New
York 2001.
112. R. Hazard and A. Tallec, Bull. Soc. Chim. Fr., 1973, 3040; M. Cariou, R. Hazard,
M. Jubault, and A. Tallec, Can. J. Chem., 61, 2359 1983.
113. W. J. Begley, J. Grimshaw, and J. Trocha-Grimskaw, J. Chem. Soc., Perkin Trans.
1, 1974, 2633; M. Dias, M. Gibson, J. Grimshaw, I. Hill, J. Trocha-Grimshaw, and
O. Hammerich, Acta Chem. Scand., 52, 549 1998.
114. K. Sasaki, H. Urata, K. Uneyama, and S. Nagura, Electrochim. Acta, 12, 137
1967.
115. V. D. Parker and B. B. Burgert, Tetrahedron Lett., 1968, 2415.
116. A. Nilsson, U. Palmquist, T. Petterson, and A. Ronlan, J. Chem. Soc., Perkin
Trans. 1, 1978, 708.
117. S. D. Ross, M. Finkelstein, and R. C. Petersen, J. Am. Chem. Soc., 88, 4657
1966; E. J. Rudd, M. Finkelstein, and S. D. Ross, J. Org. Chem., 37, 1763
1972.
118. T. Shono, Y. Matsumura, K. Tsubuta, Y. Sugihara, S. Yamane, T. Kanazawa, and T.
Aoki, J. Am. Chem. Soc., 104, 6697 1982; T. Shono, Y. Matsumura, K. Tsubata,
and K. Uchida, J. Org. Chem., 51, 2590 1986.
119. K. D. Mueller, Tetrahedron, 56, 9527 2000; H. K. Reddy, K. Chiba, Y. Sun, and
K. D. Mueller, Tetrahedron, 57, 5183 2001.
120. T. Siu, W. Li, and A. K. Yudin, J. Comb. Chem., 2, 545 2000; A. K. Yudin and
T. Siu, Curr. Opin. Chem. Biol., 5, 269 2001.
121. L. Eberson, J. Am. Chem. Soc., 89, 4669 1967; L. Eberson, Tetrahedron, 32,
2185 1976.
122. D. B. Clark, M. Fleischman, and D. Pletcher, J. Chem. Soc., Perkin Trans. 2,
1973, 1578.
123. E. Klocke, A. Matzeit, M. Gockeln, and H. J. Schaefer, Chem. Ber., 126, 1623
1993.
124. H. J. Schafer, Angew. Chem. Int. Ed. Engl., 20, 911 1981.
125. M. Feldhues and H. J. Schafer, Tetrahedron, 41, 4195, 4213 1985.
126. L. Becking and H. J. Schafer, Tetrahedron Lett., 29, 2797, 2801 1988.
127. R. N. Renaud, P. J. Champagne, and M. Savard, Can. J. Chem., 57, 2627 1979.

S32

Journal of The Electrochemical Society, 149 4 S21-S33 2002

128. N. Muller, J. Org. Chem., 48, 1370 1983: N. Muller, J. Org. Chem., 49, 2826,
4559 1984; N. Muller, J. Org. Chem., 51, 263 1986.
129. M. Yoshikawa, T. Kamigauchi, Y. Iked, and I. Kitagawa, Chem. Pharm. Bull.
(Tokyo), 29, 2582 1981.
130. D. Seebach, R. Charczuk, C. Gerber, P. Renaud, H. Berner, and H. Schneider,
Helv. Chim. Acta, 72, 419 1989.
131. D. H. Evans, P. Jimenez, and M. J. Kelly, J. Electroanal. Chem., 163, 145 1984.
132. R. A. Marcus, Annu. Rev. Phys. Chem., 15, 155 1964.
133. C. L. Wong and J. K. Kochi, J. Am. Chem. Soc., 101, 5593 1979; S. Fukuzumi,
C. L. Wong, and J. K. Kochi, J. Am. Chem. Soc., 102, 2928 1980.
134. L. Eberson, Electron Transfer Reactions in Organic Chemistry, Springer, Berlin
1987.
135. J. R. Miller, L. T. Calcaterra, and G. L. Closs, J. Am. Chem. Soc., 106, 3047
1984.
136. I. R. Gould, J. E. Moser, D. Ege, and S. Farid, J. Am. Chem. Soc., 110, 1991
1988; I. A. Gould, R. Moody, and S. Farid, J. Am. Chem. Soc., 110, 7242 1988.
137. C. P. Andrieux, J. M. Dumas-Bouchiat, and J.-M. Saveant, J. Electroanal. Chem.,
87, 39, 55 1978; C. P. Andrieux, J. M. Dumas-Bouchiat, and J.-M. Saveant, J.
Electroanal. Chem., 88, 43 1978.
138. J. W. Sease and R. C. Reed, cited in A. J. Fry, Synthetic Organic Electrochemistry, Harper & Row, 1972, p. 67; J. W. Sease and R. C. Reed, Tetrahedron
Lett., 1975, 393.
139. H. Lund, M.-A. Michel, and J. Simonet, Acta Chem. Scand., B28, 900 1974; H.
Lund, M.-A. Michel, and J. Simonet, Acta Chem. Scand., B29, 217 1975; J.
Simonet, M.-A. Michel, and H. Lund, Acta Chem. Scand., B29, 489 1975.
140. H. Lund and J. Simonet, J. Electroanal. Chem., 65, 205 1975.
141. V. G. Mairanovskii and N. F. Loginova, Zh. Obshch. Khim., 45, 2112 1975; V.
G. Mairanovskii, Angew. Chem. Int. Ed. Engl., 15, 281 1976.
142. P. E. Hansen, A. Berg, and H. Lund, Acta Chem. Scand., B30, 267 1976; E.
Hobolth and H. Lund, Acta Chem. Scand., B30, 895 1976.
143. H. Lund, K. Daasbjerg, D. Occhialini, and S. U. Pedersen, Elektrokhimiya, 31,
939 1995.
144. J. Simonet and H. Lund, Acta Chem. Scand., B31, 909 1977.
145. C. Gosden, K. P. Healy, and D. Pletcher, J. Chem. Soc. Dalton Trans., 1978, 972.
146. J. J. Pietron and R. W. Murray, J. Phys. Chem. B, 103, 4440 1999.
147. M. Brust, M. Walker, D. Bethell, D. J. Schiffrin, and R. Whyman, J. Chem. Soc.
Chem. Commun., 1994, 801.
148. R. S. Ingram and R. W. Murray, Langmuir, 14, 4115 1998.
149. A. C. Templeton, W. P. Wuelfing, and R. W. Murray, Acc. Chem. Res., 33, 27
2000.
150. W. Schuhmann, H. Zimmermann, K. Habermuller, and V. Laurinaviciuc, Faraday
Discuss., 116, 245 2000.
151. M. E. Hermandez and D. K. Newman, Cell. Med. Life Sci., 58, 1662 2001.
152. C.-H. Lee, T. A. Lewis, A. Paszczynski, and R. L. Crawford, Biochem. Biophys.
Res. Commun., 261, 562 1999.
153. W. Schmidt and E. Steckhan, Chem. Ber., 113, 577 1980; M. Platen and E.
Steckhan, Liebigs Ann. Chem., 1984, 1563.
154. T. Douadi, M. Cariou, and J. Simonet, Tetrahedron, 52, 4449 1996; T. Douadi,
M. Cariou, and J. Simonet, New J. Chem., 20, 1031 1996.
155. C. E. Lomen, W. U. de Alwis, and G. S. Wilson, J. Chem. Soc., Faraday Trans. 1,
82, 1265 1986.
156. C. Bourdillon, C. Demaille, J. Moiroux, and J.-M. Saveant, Acc. Chem. Res., 29,
529 1996.
157. N. Anicet, A. Anna, C. Bourdillon, C. Demaille, J. Moiroux, and J.-M. Saveant,
Faraday Discuss., 116, 269 2000.
158. C. P. Andrieux, I. Gallardo, and J.-M. Saveant, J. Am. Chem. Soc., 111, 1620
1989.
159. K. Daasbjerg, S. U. Pedersen, and H. Lund, in Measurements and Estimation of
Redox Potentials of Organic Radicals in General Aspects of Free Radical Chemistry, Z. B. Alfassi, Editor, p. 385, John Wiley & Sons, New York 1999.
160. D. M. Wayner and D. Griller, J. Am. Chem. Soc., 107, 7764 1985; B. A. Sim, P.
H. Milne, D. Griller, and D. D. Wayner, J. Am. Chem. Soc., 112, 6635 1990. F.
Fontana, R. J. Kolt, Y. Huang, and D. D. M. Wayner, J. Org. Chem., 59, 4671
1994.
161. V. A. Benderskii, S. D. Babenko, Y. M. Zolotovitskii, A. G. Krivenko, and T. S.
Rudenko, J. Electroanal. Chem., 56, 325 1974. V. A. Benderskii and A. G.
Krivenko, Russ. Chem. Rev., 59, 1 1990.
162. V. V. Konovalov, I. I. Bilkis, B. A. Selinavov, V. D. Sheingarts, and T. D. Tsvetkov, J. Chem. Soc., Perkin Trans. 2, 1993, 1707; P. Hapiot, V. V. Konovalov, and
J.-M. Saveant, J. Am. Chem. Soc., 117, 1428 1995, J. Gonzalez, P. Hapiot, V.
Konovalov, and J.-M. Saveant, J. Am. Chem. Soc., 120, 10171 1998.
163. R. Fuhlendorff, D. Occhialini, S. U. Pedersen, and H. Lund, Acta Chem. Scand.,
43, 803 1989; D. Occhialini, S. U. Pedersen, and H. Lund, Acta Chem. Scand.,
44, 715 1990; D. Occhialini, J. S. Kristensen, S. U. Pedersen, and H. Lund, Acta
Chem. Scand., 46, 474 1992; D. Occhialini, K. Daasbjerg, and H. Lund, Acta
Chem. Scand., 47, 1100 1993; H. Lund, K. Daasbjerg, D. Occhialini, and S. U.
Pedersen, Russ. J. Electrochem., 31, 865 1995.
164. S. U. Pedersen and T. Lund, Acta Chem. Scand., 45, 397 1991.
165. S. U. Pedersen, T. Lund, K. Daasbjerg, M. Pop, I. Fussing, and H. Lund, Acta
Chem. Scand., 52, 657 1998.
166. J. H. Wagenknecht and M. M. Baizer, J. Org. Chem., 31, 3885 1966.
167. C. Amatore, G. Capobianco, G. Farnia, G. Sandona, J.-M- Saveant, M. G. Severin, and E. Vianello, J. Am. Chem. Soc., 107, 1815 1985; M. C. Arevalo, G.
Farnia, M. G. Severin, and E. Vianello, J. Electroanal. Chem., 220, 201 1987.
168. P. E. Iversen and H. Lund, Tetrahedron Lett., 1969, 3523.
169. R. C. Hallcher and M. M. Baizer, Liebigs Ann. Chem., 1977, 737.

170. T. Shono, S. Kashimura, and H. Nogusa, J. Org. Chem., 49, 2043 1984.
171. T. Shono, N. Kise, M. Masuda, and T. Suzumoto, J. Org. Chem., 50, 2527 1985.
172. D. H. Chin, G. Chiericato, E. J. Nanni, and D. T. Sawyer, J. Am. Chem. Soc., 104,
1296 1982.
173. M. Sugawara, M. M. Baizer, W. T. Monte, R. D. Little, and U. Hess, J. Am. Chem.
Soc., B37, 509 1983.
174. Z. I. Niazimbetova, D. H. Evans, L. M. Liable-Sands, and A. L. Rheingold, J.
Electrochem. Soc., 147, 256 2000.
175. H. Sagae, M. Fujihira, H. Lund, and T. Osa, Bull. Chem. Soc. Jpn., 53, 1537
1980.
176. J. Pinson and J.-M. Saveant, J. Chem. Soc. Chem. Commun., 1974, 933; J. Pinson
and J.-M. Saveant, J. Am. Chem. Soc., 100, 1507 1978, J.-M. Saveant, Acc.
Chem. Res., 13, 323 1980.
177. J. E. Swartz and T. T. Stenzal, J. Am. Chem. Soc., 106, 2520 1984.
178. N. Alam, C. Amatore, C. Combellas, A. Thiebault, and J. N. Verpeaux, Tetrahedron Lett., 49, 6171 1987.
179. C. Costentin, P. Hapiot, M. Medebielle, and J.-M. Saveant, J. Am. Chem. Soc.,
121, 4451 1999, 122, 5623 2000.
180. T. Lund and H. Lund, Acta Chem. Scand., B40, 470 1986; T. Lund and H. Lund,
Acta Chem. Scand., B41, 93 1987; T. Lund and H. Lund, Acta Chem. Scand.,
B42, 269 1988.
181. H. Lund, K. Daasbjerg, T. Lund, and S. U. Pedersen, Acc. Chem. Res., 28, 313
1995.
182. L. Horner and D. Degner, Tetrahedron Lett., 1968, 5889; 1245 1971, Electrochim. Acta, 19, 611 1974; L. Horner and D. H. Skaletz, Liebigs Ann. Chem.,
1977, 1365.
183. D. Seebach and H. A. Oei, Angew. Chem., 87, 629 1975.
184. R. N. Gourley, J. Grimshaw, and P. G. Millar, J. Chem. Soc. Chem. Commun.,
1967, 1278 1967: R. N. Gourley, J. Grimshaw, and P. G. Millar, J. Chem. Soc.,
1970, 2318.
185. R. Hazard, S. Jaouannet, and A. Tallec, Tetrahedron, 38, 93 1982,
186. N. Schoo and H. J. Schafer, Liebigs Ann. Chem., 1993, 609; M. F. Nielsen, B.
Batanero, S. Rolvering, and H. J. Schafer, in Novel Trends in Electroorganic
Synthesis, S. Torii, Editor, Springer, Tokyo 1998.
187. J. Hermolin, J. Kopilev, and E. Gileadi, J. Electroanal. Chem., 71, 245 1976.
188. H. Wendt, Angew. Chem. Int. Ed. Engl., 21, 256 1982.
189. B. F. Watkins, J. R. Behling, E. Kariv, and L. L. Miller, J. Am. Chem. Soc., 97,
3549 1975; B. E. Firth, L. L. Miller, M. Mitani, T. Rogers, J. Lennox, and R. W.
Murray, J. Am. Chem. Soc., 98, 8271 1976.
190. S. Abe, T. Nonaka, and T. Fuchigami, J. Am. Chem. Soc., 105, 3630 1083, T.
Komoti and T. Nonaka, J. Am. Chem. Soc., 106, 2656 1984.
191. T. Osa, Y. Kashiwagi, Y. Yanagisawa, and J. M. Bobbitt, J. Chem. Soc. Chem.
Commun., 1994, 2535.
192. Y. Kashiwagi, Q. Pan, F. Kurashima, C. Kikuchi, J.-i. Anzai, and T. Osa, Chem.
Lett., 1998, 143.
193. P. N. Bartlett, D. Pletcher, and J. Zeng, J. Electrochem. Soc., 144, 3705 1997.
194. E. Steckhan, Top. Curr. Chem., 170, 83 1994; E. Steckhan, GDCh Monographie,
9, 483 1996, Gesellschaft Deutcher Chemiker, Frankfurt/Main, 1997.
195. E. Steckhan, in Electroenzymatic Synthesis in Organic Electrochemistry, 4th ed.,
H. Lund and O. Hammerich, Editors, Marcel Dekker, New York 2001.
196. B. L. Funt and K. C. Yu, J. Polym. Sci., 62, 359 1962.
197. C. Simionescu and M. Grovu, Angew. Makromol. Chem., 111, 149 1983.
198. G. Mengoli, G. Vidotto, and F. Furianetto, Makromol. Chem., 37, 203 1970.
199. G. S. Shapoval, J. Macromol. Sci., Chem., A17, 453 1982.
200. H. Shirakawa, E. J. Louis, A. G. MacDiarmid, C. K. Chiang, and A. J. Heeger, J.
Chem. Soc. Chem. Commun., 1977, 578 1977.
201. P. J. Nigrey, A. G. MacDiarmid, and A. J. Heeger, J. Chem. Soc. Chem. Commun.,
1979, 594.
202. A. F. Diaz, K. K. Kanazawa, and G. P. Gardini, J. Chem. Soc. Chem. Commun.,
1979, 635.
203. G. Tourillon and F. Garnier, J. Electroanal. Chem., 135, 173 1982.
204. A. F. Diaz and J. A. Logan, J. Electroanal. Chem., 111, 111 1980.
205. C. P. Andrieux, P. Audebert, P. Hapiot, and J.-M. Saveant, J. Am. Chem. Soc., 112,
2439 1990.
206. P. Audebert, P. Hapiot, J.-M. Pernaut, and P. Garcia, J. Electroanal. Chem., 361,
283 1993.
207. A. Smie, A. Synowczyk, J. Heinze, R. Alle, P. Tschuncky, G. Gotz, and P. Bauerle,
J. Electroanal. Chem., 452, 87 1998; J. Heinze, H. John, M. Dietrich, and P.
Tschuncky, Synth. Met., 119, 49 2001.
208. M. A. Vorotyntsev and J. Heinze, Electrochim. Acta, 46, 3309 2001.
209. E. M. Genies, G. Bidan, and A. F. Diaz, J. Electroanal. Chem., 149, 101 1983.
210. T. Kobayashi, H. Yoneyama, and H. Tamura, J. Electroanal. Chem., 161, 419
1984, 177, 281 1984.
211. A. Kitani, J. Yano, and K. Sasaki, J. Electroanal. Chem., 209, 227 1986.
212. T. Kobayashi, H. Yoneyama, and H. Tamura, J. Electroanal. Chem., 161, 419
1984.
213. E. Meggers, M. E. Michel-Beyerly, and B. Giese, J. Am. Chem. Soc., 120, 12950
1998, E. Meggers, D. Kusch, M. Spichty, U. Wille, and B. Giese, Angew. Chem.
Int. Ed. Engl., 37, 460 1998.
214. B. Hall, R. E. Holmlin, and J. K. Barton, Nature (London), 382, 731 1996.
215. D. Ly, L. Sanii, and G. B. Schuster, J. Am. Chem. Soc., 121, 9400 1999.
216. K. Fukui and K. Tanaka, Angew. Chem. Int. Ed. Engl., 37, 158 1998.
217. H. Sugiyama and I. Saito, J. Am. Chem. Soc., 118, 7063 1996.
218. H.-W. Fink and C. Schonenberger, Nature (London), 398, 407 1999.
219. M. A. Brett, A. M. O. Brett, and S. H. P. Serrano, J. Electroanal. Chem., 366, 225
1994.

Journal of The Electrochemical Society, 149 4 S21-S33 2002


220. M. Faraggi, F. Broitman, J. B. Trent, and M. H. Klapper, J. Phys. Chem., 100,
1475 1996.
221. D. L. Croteau and V. A. Bohr, J. Biol. Chem., 272, 25409 1997.
222. A. Heller, Faraday Discuss., 116, 1 2000.
223. L. Walder, in Organic Electrochemistry, 3rd ed., H. Lund and M. M. Baizer,
Editors, p. 809, Marcel Dekker, New York 1991.
224. J.-i. Yoshida and S. Suga, in Organic Electrochemistry, 4th ed., H. Lund and O.
Hammerich, Editors, p. 765, Marcel Dekker, New York 2001.
225. W. J. Settineri and L. D. McKeever, in Technique of Electroorganic Synthesis, Part
II, p. 397, John Wiley & Sons, New York 1975.
226. G. Gritzner and J. Kuta, Pure Appl. Chem., 56, 461 1984; G. Gritzner and J.
Kuta, Electrochim. Acta, 29, 869 1984.
227. W. Eisenbach and H. Lemkuhl, Dechema-Monogr., 98, 269 1985.
228. K. M. Kadish, in Progress in Inorganic Chemistry, Vol. 34, S. J. Lippard, Editor,
p. 435, John Wiley & Sons, New York 1986.
229. D. Lexa and J.-M. Saveant, J. Am. Chem. Soc., 100, 3220 1978; D. Lexa and
J.-M. Saveant, Acc. Chem. Res., 16, 235 1983.
230. R. Scheffold, Chimia, 39, 203 1985, R. Scheffold, Nachr. Chem., Tech. Lab., 36,
261 1988; S. Abrecht and R. Scheffold, Chimia, 39, 211 1985.
231. S. Ozaki, H. Matsushita, and H. Ohmori, J. Chem. Soc., Perkin Trans. 1, 1993,
649.
232. S. Torii, H. Tanaka, and K. Morisaki, Tetrahedron Lett., 26, 1655 1985.

S33

233. M. Troupel, Y. Rollin, S. Sibille, J.-P. Fauvarque, and J. Perichon, Chem. Res.
Synop., 24, 26 1980.
234. T. Itoh, S. Emoto, M. Kondo, H. Ohara, H. Tanaka, and S. Torii, Electrochim.
Acta, 42, 2133 1997.
235. J.-i. Yoshida, T. Murata, and S. Isoe, Tetrahedron Lett., 27, 3373 1986.
236. T. Shono, Y. Matsumura, S. Katoh, and N. Kise, Chem. Lett., 1985, 463.
237. J.-i. Yoshida, K. Muraki, H. Funahashi, and N. Kawabata, J. Org. Chem., 51, 3996
1986.
238. J.-i. Yoshida, M. Watanabe, H. Toshioka, M. Imagawa, and S. Suga, in Novel
Trends in Electroorganic Synthesis, S. Torii, Editor, p. 99, Springer, Berlin 1998;
J.-i. Yoshida, M. Watanabe, H. Toshioka, M. Imagawa, and S. Suga, J. Electroanal. Chem., 507, 55 2001.
239. L. L. Bolt, Hydrocarbon Process, 44, 115 1965.
240. M. M. Baizer, J. Electrochem. Soc., 111, 215 1964.
241. J. Chaussard, M. Troupel, Y. Robin, C. Jacob, and J. P. Jubasz, J. Appl. Electrochem., 19, 345 1989.
242. A. Fontana, Chim. Nouv., 11, 1232 1993.
243. M. M. Baizer, in Organic Electrochemistry, 3rd ed., H. Lund and M. M. Baizer,
Editors, p. 1421, Marcel Dekker, New York 1991.
244. H. Putter and H. Hannebaum, in New Directions in Organic Electrochemistry, A.
J. Fry and Y. Matsumura, Editors, PV 2000-15, p. 25, The Electrochemical Society Proceedings Series, Pennington, NJ 2001.

You might also like