Sei sulla pagina 1di 19

k

GENERAL ASPECTS OF REDOX


CHEMISTRY

Felipe J. González
Departamento de Química del Centro de Investigación y Estudios Avanzados del I.P.N, México City,
México

Carlos Frontana
Centro de Investigación y Desarrollo Tecnológico en Electroquímica, Parque Tecnológico
Querétaro Sanfandila, Sanfandila, Pedro Escobedo, Querétaro, México

Martín Gómez
Departamento de Sistemas Biológicos, Universidad Autónoma Metropolitana, Unidad Xochimilco,
México City, México
k Ignacio González k
Departamento de Química, Universidad Autónoma Metropolitana-Iztapalapa, San Rafael Atlixco,
México City, México

1 INTRODUCTION

1.1 Redox Chemistry


Chemical reactions are permanently occurring in almost all systems known and they are
activated in different ways. Thermal and photochemical activation modes are the most
common, which are of great relevance for synthetic purposes. The third mode involving
the electronic exchange between molecules is of interest too, both as synthetic tool and to
describe many chemical and biochemical processes. This type of molecular activation gives
rise to redox chemistry in which an oxidizing and a reducing species are transformed. In
the general case of a redox reaction, one electron is transferred from the highest occupied
molecular orbital (HOMO) of the reducing species toward the lowest unoccupied molecular
orbital (LUMO) of the oxidizing species, generating, in many cases, paramagnetic interme-
diates, such as radical cations, radical anions, and free radicals, which have a higher level

Encyclopedia of Physical Organic Chemistry, First Edition. Edited by Zerong Wang.


© 2017 John Wiley & Sons, Inc. ISBN 978-1-118-46858-6.

k
k

2 ENCYCLOPEDIA OF PHYSICAL ORGANIC CHEMISTRY

of energy than precursor molecules, thereby presenting higher reactivity. The energies of
HOMO and LUMO orbitals are directly related to the ionization potential and electron affin-
ity parameters of the reducing and oxidizing species and have specific values dependent on
the molecular structure. Redox chemistry can thus be viewed as an interaction between a
substrate and a molecular electrode with a constant potential. Owing to this, mechanistic
aspects of redox reactions have been better analyzed in the framework of electrochemical
techniques, which allow tuning the electron transfer (ET) reaction by controlling the elec-
trode potential. Despite this advantage of electrochemistry, redox chemistry turns out to
be more versatile for synthetic purposes because it does not require the use of electrolytes
and because high concentrations of substrate can be transformed in relatively short reaction
times. Under this perspective, redox chemistry can be approached from the point of view of
mechanistic analysis with electrochemical tools or from a synthetic context with qualitative
scales of the oxidizing/reducing power of different compounds.

1.2 Outer and Inner Sphere Electron Transfers


Redox reactions can occur through two different mechanisms, which were originally for-
mulated in the field of inorganic chemistry at the beginning of 1950s [1]. A redox reac-
tion following an outer-sphere mechanism is an intermolecular ET process in which the
electronic interaction between reactants in the transition state is inexistent or very weak
(<16 kJ mol−1 ). In this mechanism, the solvation sphere of the reactant and the product are
essentially the same before and after the ET reaction. Such condition is generally satisfied
when the size of the molecule is relatively large and when the charge formed in the oxida-
tion or reduction state is adequately stabilized by electronic or resonance effects. Under this
k context, a simple example is found in the self-exchange reactions, which have been used to k
construct the modern theories of ET. In these isoergonic reactions, the driving force is zero
(ΔG∘ = 0) and there is no net chemical change, as illustrated in Equation 1 [2].

[Ru(bpy)3 ]3+ + [Ru(bpy)3 ]2+ = [Ru(bpy)3 ]2+ + [Ru(bpy)3 ]3+ (1)

This category of reactions includes isotopic self-exchange reactions, for example,


between Fe2+ and the isotopically tagged Fe*3+ ions, which also allow the elimination of
the thermodynamic stability factor between reactants and products [3].

Fe2+ + Fe∗3+ = Fe3+ + Fe∗2+ (2)

A wide variety of organometallic compounds have also been recognized as outer sphere
ET species, being the typical example, the ferrocene, which is reversibly transformed into
the ferrocenium ion. In organic chemistry, great molecules, such as fused aromatic hydro-
carbons, quinones, and nitroaromatic compounds, also present an outer sphere behavior
when they are reduced in aprotic conditions to yield relatively stable radical anions.
On the other hand, the inner sphere ET processes involve a close interaction between
reactants resulting in the formation of an adduct that allows the ET and ligand transference.
In this adduct, a ligand is the bridge between the reactants. A well-recognized inner sphere
ET system was proposed by Taube for the next reaction in acidic medium:

5H2 O + [Cr(OH2 )6 ]2+ + [Co(NH3 )5 Cl]2+ → adduct →


[Cr(OH2 )5 Cl]2+ + [Co(OH2 )6 ]2+ + 5NH3 (3)

k
k

GENERAL ASPECTS OF REDOX CHEMISTRY 3

Here, the ET and the bond breaking/formation reactions are carried out; however, the mixed
valence metal complexes present in the mentioned adduct, are a simple class of inner sphere
ET process, in which the intramolecular ET occurs without any modification of the chemical
nature of bonds [4].
When this sort of reactions are viewed from the electrochemical point of view, the inner
sphere mechanism can involve the adsorption of either the reactant or the product as well
as processes in which the ET reaction is concerted with a bond breaking reaction [5].

1.3 Compounds with Redox Behavior


It is well recognized that organic and inorganic species can be forced to uptake or release
electrons in electrochemical experiments over a wide range of electrode potentials, which
implies in principle that any substrate could work as an oxidizing or reducing species
under appropriate conditions. However, in the preparative redox chemistry practice, the
most commonly used oxidizing and reducing compounds are those whose redox potential
is rather inside a limited and accessible solvent–electrolyte electroactivity window. In this
framework, a wide variety of neutral, anionic, and cationic species (used as typical redox
compounds) can be characterized by electrochemical methods.
Many redox species can be obtained from the oxidation or reduction of neutral
molecules, giving rise to anionic and cationic organic compounds. A single general
example that illustrates this issue starts from a molecule (Mn ) that can be transformed by
successive ET steps into the species (Mn−1 ) and (Mn−2 ). Due to the fact that the redox
potential of the couple M/Mn−1 is less negative than the redox potential of Mn−1 /Mnn−2 ,
the reducing and oxidizing power for each one of these species can be established on an
k arbitrary scale, as depicted below. k

Oxidizing power

+e– +e–
Mn Mn – 1 Mn – 2 n = 0, 1, 2,...
–e– –e–

Reducing power

According to this scheme, the reducing power increases with the negative charge on the
species while the oxidizing power in this series increases in the opposite direction. It means,
for example, that radical anions and dianions are better reducing agents than the respective
neutral derivatives and free radicals. A similar argumentation can be deduced for species
resulting from the successive uptake of electrons, as shown below.

Oxidizing power

–e– –e–
Mn Mn + 1 Mn + 2 n = 0, 1, 2,...
+e– +e–

Reducing power

By analogy to the previous case, the oxidizing power increases with the positive charge
on the species, while the reducing power in this series increases in the opposite direction.

k
k

4 ENCYCLOPEDIA OF PHYSICAL ORGANIC CHEMISTRY

That is, positively charged species, such as radical cations and carbocations, are better oxi-
dizing species than the corresponding neutral derivative and free radicals.
In conclusion, the species with higher reducing power are those possessing more nega-
tive redox potentials, while more oxidizing species present less negative or more positive
values. This thermodynamically based argumentation is related to the redox potential of the
reducing or oxidizing species; however, when ionic species are considered as reductants or
oxidants, it is relevant to consider also the intervention of ion pairing effects, which can
decrease the rate of electron exchange of the redox reaction. A wider revision of the pos-
sibilities of different compounds to react as oxidants and reductants is available elsewhere
and a summary is given in the following two sections [6].

1.3.1 Inorganic and Organic Oxidants Organic oxidant species represent a wide variety
of compounds that have been used to perform organic and inorganic chemical reactions.
Due to the properties of metals to easily change their oxidation state by ET, the number of
inorganic oxidants available is greater than that of organic.
The case of ferricenium ions is particularly relevant due to the fact that the redox poten-
tial can be tuned by appropriate substitutions on the cyclopentadienyl rings and that this ion
can be used in different solvents following outer-sphere ET mechanisms. Silver salts belong
to other relevant category of oxidizing agents together with halogens, copper salts, FeCl3 ,
the Ce4+ ion, metal anionic complexes such as [IrCl6 ]2− and [PtCl6 ]2− , neutral complexes
such as dithiolene derivatives Ni(tfd)2 , transition metal hexahalides (MoF6 , WF6 , and UF6 ),
alkyl halides, arsenic and antimony pentahalides, and AlCl3 , among others.
Organic oxidants are less common, except for nitrosonium [NO]+ and aryldiazonium
salts [N2 Ar]+ , whose use is widely recognized. This group also includes arylaminium radi-
k cal cations [N(Ar)3 ]+ , which are prepared by the oxidation of triarylamines in the presence k
of a wide variety of counter ions, such as PF6 − , BF4 − , SbCl6 − , or WCl6 − . As in the case
of ferricenium derivatives, the oxidizing power of arylaminium cations can be modulated
by an adequate substitution of aryl groups with electron-releasing or electron-withdrawing
substituents, which allow the modification of the redox potential over a wide potential range
(i.e., from 0.16 to 1.72 V vs Fc). Other radical cations that have been used as organic oxi-
dants have been synthesized via reaction with nitrosonium salts. Examples of these redox
intermediates are the heterocyclic compounds derived from thianthrene, phenoxathiin, and
phenothiazine, which have been prepared with a wide variety of counteranions.
Carbocations, such as triphenylcarbenium cations [Ph3 C]+ , are more reactive species
than radical cations and their redox potential can be regulated through ring substitutions.
These derivatives are considered as mild oxidants that are reduced to the corresponding
triphenylmethyl radicals, which are considered to be some of the most stable benzylic rad-
icals. To stabilize these mild oxidants, several anions such as BF4 − and PF6 − have been
used and in the simplest case, concerning the nonsubstituted intermediate, the commercial
source salts. Weaker oxidant cations are tropylium ions [C7 H7 ]+ , whose redox potential
can also be tuned by ring substitution.
Organic compounds that give rise to radical anions through ET have also been used as
oxidants. Examples are found in the group of alkyl halides, percyanocarbons, such as tetra-
cyanoquinodimethane and quinones. In the case of alkyl halides, the ET reaction releases
free radicals while in the other cases, stable radical anions and dianions can be formed.
However, in quinone compounds, the oxidant power can be modulated either by substi-
tution in the quinonoid moiety or by the presence of proton donors of different acidity
levels.

k
k

GENERAL ASPECTS OF REDOX CHEMISTRY 5

1.3.2 Inorganic and Organic Reductants Due to the difficulty to reduce alkali cations
at very negative potentials, the group of alkali metals has been extensively used as strong
reductants, although some limitations about fire hazards and the formation of oxides on
the surface of these metals have been recognized. Because of the huge differences between
the solvation energies between the neutral and cationic species, the oxidant power of these
metals is quite sensitive to the solvent nature, which must be mainly aprotic, and free of
water traces. In the same line, the reducing power of metal hydrides, such as NaH, NaBH,
LiAlH4 , and alkyl–lithium compounds, have also been recognized. By analogy with the
ferricenium derivatives as oxidants, organometallic compounds such as cobaltocene have
been used as reductants, being its reductant strength tuned by appropriate substitutions in
the cyclopentadienyl rings.
Organic reducing species are mainly ionic and they can be formed in different ways.
The naphthalene radical anion can be formed electrochemically in aprotic conditions at
very high reduction potentials or synthesized under inert atmosphere in dry tetrahydrofu-
rane by mixing stoichiometric amounts of naphthalene and an alkali metal. Due to the high
basicity of the naphthalenide ion, its reducing strength is sensitive to ion pairing effects
with the metal cation and solvent. Another organic reductant is the benzophenone radi-
cal anion which can also be prepared by reaction with alkali metals. Both anionic species
have very high negative redox potentials and are sensitive to the presence of oxygen in the
medium.

1.4 Physicochemical Parameters Determining a Redox Reaction

k A redox process can be viewed as a homogeneous ET reaction between a substrate S and k


either an oxidizing (Ox) or a reducing compound (Red). For example, in the case of the
reduction of S, this substrate is transformed to S− , while a reducing species (Red) is trans-
formed into the corresponding oxidizing species (Ox) (Equation 4).

S + Red ⇌ S− + Ox (4)

A quantitative approach to the extension in which reaction 4 occurs can be obtained


from two separated electrochemical reactions, both established from the point of view of
reduction processes as convention, that is, the electrochemical reduction of S and the elec-
trochemical reduction of Ox (Equations 5 and 6).

S + e− ⇌ SH− (5)

Ox + e− ⇌ Red (6)

In each case, the redox potential, described by the Nernst equation, is the physicochem-
ical parameter of interest, from which the next relationships are obtained (Equations 7
and 8). ( )
RT [Ox][S− ] RT
E∘S∕S− − E∘Ox∕Red = ln = ln Kredox (7)
F [Red][S] F

( )
F
Kredox = exp E∘S∕S− − E∘Ox∕Red (8)
RT

k
k

6 ENCYCLOPEDIA OF PHYSICAL ORGANIC CHEMISTRY

where E∘S∕S− and EOx∕Red


∘ represent formal redox potentials of the corresponding redox
couples (Equations 5 and 6) and Kredox is the equilibrium constant, associated with the
standard Gibbs free energy of redox reaction 4.
Equation 8 shows that when the redox potential of substrate E∘S∕S− is more negative

than the redox potential EOx∕Red , the value of Kredox is lower than the unity and, therefore,
reaction 4 will not proceed to an appreciable extension. This is a necessary condition to
state that the ET is not favorable; however, if the species S− is reactive enough to promote
an alternative chemical reaction possessing a quite negative free energy, then the ET step
as well as the overall process will be promoted.
On the other hand, when the redox potential of the reducing species EOx∕Red ∘ is more

negative than the redox potential ES∕S− , the value of Kredox is higher than the unity and
therefore reaction 4 will proceed spontaneously. This reasoning was exemplified for the
reduction of a substrate S; however, a conclusion for its oxidation is straightforward.
This approach is essentially thermodynamic; however, the occurrence of the homoge-
neous ET can also be limited by the ET kinetics. Therefore, a microscopic analysis of the
factors determining the values of the redox potential as well as the kinetic properties is
discussed later using the microscopic theories of ET.

1.5 The Role of Electrochemistry in the Study of Redox Reactions


The electron exchange between two species in solution is mainly determined by relative
energy differences between the HOMO and LUMO of the reductant and oxidant respec-
tively. These energies can be viewed, respectively, in relation to the ionization potential and
k electron affinity parameters, which otherwise are related to redox potentials of each one of k
these species. Under this perspective, the knowledge of the values of these potentials is
convenient to evaluate the reliability of the electron exchange between the reductant and
the oxidant. This is a first approach to the problem and it works well in many practical sit-
uations. For example, the reduction of a substrate (S) can be performed spontaneously by a
reductant (Red), when both redox potentials satisfy the condition EOx∕Red ≪ ES∕S− , while in
the case of the oxidation by an oxidant (Ox′ ), the required condition is ERed′ ∕Ox′ ≫ ES∕S+ .
In both cases, such relationships give rise to negative values for the Gibbs free energy
of the homogeneous electron exchange process. If such relationships were opposed, the
redox processes would be unfavorable; however, the intervention of coupled chemical reac-
tions, consecutive to formation of either S− or S+ , could be the key factor contributing to
the overall Gibbs free energy and therefore the factor driving the activation by electron
exchange.
Electrochemical methods are recognized as the best tool to analyze mechanistic aspects
of redox reactions, where the working electrode assumes the function of either reductant
or oxidant species. A wide diversity of methods is available; however, cyclic voltamme-
try is the most used method [7]. Today, the instrumentation does not represent a problem
for obtaining current potential curves that can later be rigorously analyzed using all the-
oretical protocols developed for this technique in the final decades of the twentieth cen-
tury. In this way, single reversible ETs for reductants and oxidants can be easily induced
and analyzed. These chemical species can be reduced or oxidized to yield highly reactive
intermediates that participate in coupled chemical reactions, which frequently provoke the
intervention of subsequent ET reactions. A part or the total sequence of these reactions can
be established from cyclic voltammetry experiments and, in some cases, using additional

k
k

GENERAL ASPECTS OF REDOX CHEMISTRY 7

techniques such as coulometry, preparative electrolysis, chronoamperometry, and station-


ary state techniques. The electrochemical methods are clearly advantageous in revealing
mechanistic aspects of the processes activated by ET; however, for preparative purposes,
the synthetic electrochemistry is time consuming, it is generally performed with low con-
centrations of substrate and requires the separation of supporting electrolyte. Alternatively,
the use of reductants and oxidants is preferred to synthesize products at the macroscale.

2 GENERAL ASPECTS OF ELECTRON TRANSFER IN ORGANIC SYSTEMS

ET processes play a predominant role in organic systems as a factor that can determine
a reaction rate. For example, in addition or elimination reactions, such as aromatic nitra-
tion, nucleophilic substitution, and organometallic additions[8–10], the intermediates gen-
erated by ET reactions are involved in the rate-determining step due to their radical and/or
nucleophilic–electrophilic character.
However, in organic compounds, ET processes cannot be treated in the same general
manner as in inorganic compounds, where the uptake of one or many electrons is directly
related to changes in the oxidation number of metal centers [11, 12]. For organic com-
pounds, an oxidation process is mostly understood as the conversion of a functional group
in a molecule from one category to another, and vice versa [13]; therefore, it is conve-
nient to compare the changes in the number of hydrogen and oxygen atoms accompanying
a given reaction. Unfortunately, this particular definition includes reactions that can be
mechanistically understood by other general pathways and not necessarily as ET processes,
for example, in the SN 2 reaction for converting bromomethane into methanol by adding
k water (which would be classified as an oxidation due to the oxygen gaining) or in bro- k
momethane conversion to methane by adding LiAlH4 (a reduction step due to the hydrogen
gaining). In addition, changes in bond order in elimination reactions, like the passage from
R1 (R2 )–CH–CH–R1 (R2 ) to R1 (R2 )–CH CH–R1 (R2 ), are not classically considered oxida-
tion processes. This problem in pathway assignment – strongly dependent on the agent
required for the transformation– makes difficult to classify ET reactions in organic materi-
als. Wiberg [14] considered a general classification, which can be enumerated as follows:

i. Direct electron transfers.


In these reactions, there is an effective change in the number of electrons between
the reactants, for example, the Kolbe reaction or the Birch reduction processes.
ii. Hydride transfers.
These mechanisms involve a direct hydride addition into a molecule, for example,
in the Cannizzaro reduction.
iii. Hydrogen atom transfers.
Reactions falling into this classification typically involve propagation of a radical
species, where a carbon center radical is effectively generated.
iv. Formation of ester intermediates.
In this kind of reaction, an elimination process typically occurs due to the sta-
bilization of an ester-like intermediate upon adding an inorganic Lewis acid (i.e.,
MnO3 ).
v. Displacement mechanisms.
Here, the organic compound acts as a displacement agent for the electronic charge
in a given electrophile.

k
k

8 ENCYCLOPEDIA OF PHYSICAL ORGANIC CHEMISTRY

vi. Addition–elimination mechanisms.


These complex mechanisms normally involve an effective addition of the reduc-
ing/oxidizing agent (i.e., to a double bond), followed by rearrangement and, finally,
elimination of a fraction of the added compound, which presents a change in the
oxidation number with respect to the original.

Even though this classification is often used to distinguish between reactive systems for
organic compounds, there is a strong dependency on specific pathways associated with the
chemical nature of both reactive species and therefore, a general approach based on these
structures would be desirable. In terms of inorganic ions, ET processes can be ordered in
two general schemes: inner and outer sphere reactions, in which the analysis is centered on
whether (or not) the coordination sphere of the reactant ions changes during the ET process.
For organic compounds, such generalization is not straightforward, because the ET occur-
ring should be considered as a nonadiabatic pathway (lacking of bond breaking/formation
processes in the transition state), compared with the classical adiabatic organic reactions.
Litter [15, 16] considered that instead of using the “sphere” classification, organic reac-
tions can be considered as occurring between “nonbonded” and “bonded” reactants (or
outer and inner sphere, respectively, for inorganic ions) and so, the classification of organic
ET reactions can be reordered as follows:

• Nonbonded (adiabatic) routes: ET and hydrogen atom transfers.


• Bonded (adiabatic) routes: Oxidation/reduction of organic ligands in metal complexes
and atom/group transfers.

k It should be regarded that such classifications apply only to those very specific cases k
where the ET is the only step occurring. In organic chemistry, it is more convenient to
regard both steps as catalytic in which the addition of an electron (or its removal) determines
a synthetically valuable route. Therefore, a few aspects should be discussed, considering
how structural factors determine the reactivity in systems where ET is occurring.

2.1 Description of Reactivity for Organic Electron Transfer Reactions


Analysis of reactive trends for “simple” adiabatic nonbonded ET systems is the topic to be
discussed. Empirically, reactive trends are known to be systematically described by nonlin-
ear free energy relationships, defined by the classical Hammett relationship:
( )
K
log = 𝜎𝜌 (9)
K0

In this equation, the changes in free energy are directly related to an empirical parameter
(𝜎), which accounts for the electronic influence of the substituent in the series of molecules
presenting a similar reaction sequence. For nonbonded ET processes, Zuman has exten-
sively described how this equation can be applied to predict oxidation or reduction energies
(in terms of their normal redox potential, E∘ ) by applying the Hammett–Zuman relation-
ship [17].
ΔE∘ = 𝜎𝜌 (10)

Even though this approach has been successfully generalized for many simple reactive
systems [18–22], it has an inherent limitation: the empirical nature of the 𝜎 values that

k
k

GENERAL ASPECTS OF REDOX CHEMISTRY 9

requires amassing large numbers of data considering the presence of many substituents
under different chemical environments. To avoid this, theoretical quantum chemical meth-
ods have provided a valuable and alternative tool for analyzing and predicting reactive
trends for series of compounds. For the special case of ET reactions and from the molecu-
lar orbital theory, it is considered that reactivity is mainly determined by the energy of the
electrons occupying the highest available electronic levels (frontier orbitals). For a general
reaction between an electron acceptor (oxidizing agent, O) and an electron donor (reducing
agent, R), a general sequence involves an effective transfer between electrons occupying
the HOMO in the R species, toward the less-energetic and available orbital in the O species,
referred as LUMO. Considering that this route should maintain adiabatic effects related to
energetic requirements for “matching” the energy of the systems, such as vibrational exci-
tations and reorganization of the inner spheres, only very small bond distortions must be
considered. Although these considerations are not expected to occur for many systems,
there are a lot of reports where the energy of organic ET systems is directly related to the
energies of the frontier orbitals, for example, by comparing ultrahigh vacuum spectroscopy
data (from which HOMO energies can be estimated), with electrode potentials for a series
of organic semiconductors [23], or considering quantum chemical calculations [24–27].
Within the framework of Marcus theory [28], the most important factor determining the
rate of ET in the studied systems is the solvent reorganization, as is expected, due to the
minimum changes occurring in internal coordinates.
Another alternative is provided by analyzing the changes in electron density in the sys-
tems, considering the analogy between the oxidation and reduction process with the respec-
tive ionization potential and electron affinity for a specific molecule. In this context, the
calculation of the changes in Gibbs free energy has proved more useful than the frontier
k orbitals for analyzing reactive trends [29–31]. k

2.2 Solvent Effects on Electron Transfer Reactions


Solvent effects on ET processes for organic reactions are scarcely considered, compared to
the classical descriptions for inorganic compounds [32]. In particular, the nature of the sol-
vent is especially considered when the charge of the reactive intermediate is of importance
for further steps, for example, the electrogenerated anion of tetramethylporphyrin is stabi-
lized upon increasing the polarity of the solution, as exemplified upon studying its stability
in DMSO/water mixtures [33]. In addition, DMSO solutions have provided carbanion sta-
bilization long enough to study quenching effects via photochemistry [34]. An interesting
example considering the chemical changes between carbocation/carbanion chemistry into a
radical one was presented by Arnett [35], where addition of bis(4-dimethyl-aminophenyl)
phenylcarbenium tetrafluoroborate to sodium tris(4-nitrophenyl) methanide leads to cor-
responding triphenylmethyl radicals via a one-electron process, when tetrahydrofuran is
employed. The same reaction performed in a more polar solvent (sulfolane) leads to stabi-
lization of the charged species, thus avoiding radical formation. The equilibrium between
the charged participant and the radical species can be systematically analyzed by evaluation
of the corresponding UV–vis spectra in mixtures of these solvents [35]. Disproportionation
of the 1-ethyl-4-(methoxycarbonyl)pyrinidyl radical into an ion pair has also proven to be
solvent dependent, where the increase of the solvent polarity leads to a better stabilization
of the ion pair, thus transforming the radical [36].
Kinetically, solvent effects have been considered to play a significant role in the rates
of ET [37]. However, solvent effects cannot be considered for species that involve large

k
k

10 ENCYCLOPEDIA OF PHYSICAL ORGANIC CHEMISTRY

changes in bond energy or considerable conformational energies in the transition state


[38]. Normally, the reorganization energies are calculated, as a fair approach, as if they
are spheres [39]. In terms of the full use of Marcus theory, organic compounds have proven
very useful, because rather than inorganic ions, they have been considered for extensive
calculations of the relative amount of reorganization energies and for tests of the nature
of the Marcus inverted region; this was the case of the studies of fluorescence quenching
kinetics of excited stated by Rehm and Weller [40]. In general, it is found that larger values
of solvation shells tend to increase the values of the reorganization energy thus generating
slow ET.

3 REDUCTION AND OXIDATION PROCESSES

3.1 Reduction Processes of Organic Compounds


Details of ET reduction of organic compounds have been mainly analyzed by electrochem-
ical methods [41]. In this case, the molecular activation is induced by the electrode instead
of a reductant, however, once the first reduced species are formed, it follows a pathway that
is intrinsic of the chemical properties of the intermediates, such as the one occurring in the
traditional synthetic redox chemistry. Depending on the molecular structure, several func-
tional organic groups can be reduced under different conditions to yield different products,
however, only a brief survey will be presented in the following paragraphs.
Hydrocarbons can be classified into two general groups, saturated and unsaturated. The
first group is difficult to reduce under typical electrolysis conditions due to their very low
k electron affinity; however, the presence of double or triple bonds in the structure makes the k
reduction process possible. In the case of aromatic hydrocarbons (A) with several fused
rings, the reduction can be carried out in two steps, giving rise to relatively stable radical
anions (A•− ) and reactive dianions (A2− ). In the presence of proton donors, these species
are reactive enough to be protonated, yielding the corresponding dihydro derivatives (AH2 )
[42].
+e− +H+ +e− +H+
−−−−−−⇀
A↽ −− A•− −
↽ ⇀
−−−−−− −−−−−⇀
− AH• −
↽ −− AH− −
↽ ⇀
−−−−−−− AH2

Due to the presence of species with radical character, polymerization processes are pos-
sible [43]; however, when reduction is performed in the presence of alkyl halides and a
proton source, monoalkylation of the hydrocarbon also takes place [44].
Halogenated organic compounds can be reduced to unstable radical anions, which can
be further decomposed into a free radical by halide loss [45]. In the simplest case of an
alkyl halide RX, the reduction is carried out at high potentials and gives rise to the radical
anion RX•− , which is cleaved into X− and the free radical R• . This intermediate can react
in solution or be reduced to a carbanion R− . In this way, products of radical coupling or
reactions with electrophiles are possible.

+e– – –X– –
RX RX R R Carbanion chemistry

Radical chemistry

k
k

GENERAL ASPECTS OF REDOX CHEMISTRY 11

Although different structures have been tested to obtain a wide variety of products, this
reaction is essentially based on the typical reactions of free radicals and carbanions, which
are sensitive to reaction conditions.
Reduction of nitro and related compounds involves a complicated and rich chemistry
that depends on the acidity level of the medium and the electrolysis potential. In some
cases, nitroso and hydroxylamines are intermediate products in the passage to amine-type
compounds. To illustrate the advantages of the redox and electrochemical activation of this
sort of compounds, the case of aromatic nitro compounds (ArNO2 ) was taken.

+ 2e– + 2H+ + 2e– + 2H+ + 2e– + 2H+


ArNO2 ArNO ArNHOH ArNH2
– H2O – H2O

Depending on the electrolysis potential or the strength of some reductants present in the
acidic medium, the direct transformation into the hydroxylamine (ArNHOH) or the amine
compound (ArNH2 ) can be driven, respectively, by exchanging four or six electrons.
The group of carbonyl compounds is diverse, but ketones and aldehydes can be mainly
found here. Owing to the electrophore character of the carbonyl group, most of the com-
pounds containing this functionality can be reduced either by a reductant or by an electrode.
In the absence of any other functional group, the first step of the mechanism involves for-
mation of a radical anion [46], whose reactivity against proton donors is increased from
aromatic to aliphatic carbonyl compounds. Depending on the experimental conditions and
the molecular structure, this intermediate can be observed by electron spin resonance spec-
troscopy; however, in the presence of a proton donor, a protonated radical is produced. This
k radical is more easily reduced than the original molecule and a second electron and proton k
transfer occurs to finally yield the corresponding alcohol [47, 48].

R′ R′ R′
+e– R′ R′
+H+ +e– _ +H+
O O – OH OH H OH
R R R R
R

R′
OH OH R′
R
R R

R′ OH

In competition with the protonation sequence yielding the alcohol, the dimerization of
the protonated radical to afford the pinacol structure is also possible. A wide variety of
alternative reactions of the radical species have been observed, for example, reactions in
the presence of alkenes and carbon dioxide to produce, respectively, carbon–carbon bonds
[49] and carboxyl groups [50]. Among other carbonyl-type compounds, the reduction of
quinones has been considered as very important due to their function as electron and proton
carriers in biological processes [51].
Not less relevant examples of reduction of organic compounds are also available, for
example, the reduction of fullerene, ammonia, hydrazine, oxime, carboxylic acid, ester,
anhydride, acyl halide, and nitrile derivatives. A detailed description of these processes is
out of the scope of this chapter; however, an extensive revision can be found in the litera-
ture [41].

k
k

12 ENCYCLOPEDIA OF PHYSICAL ORGANIC CHEMISTRY

3.2 Oxidation Processes of Organic Compounds


The oxidation processes of organic compounds are also so vast as the reduction ones and in
some cases, the compounds that can be reduced, can also be oxidized. The oxidation by an
oxidant or an electrode is performed by electron abstraction from the HOMO orbital. The
oxidation process is generally performed on aromatic and unsaturated hydrocarbons (RH)
[41]. Depending on the molecular charge, radical cations (RH•+ ) or free radicals (R• ) can
be obtained. In the case of radical cations, these are highly reactive intermediates that can
react easily with both nucleophilic species (i.e., Nu− ) or basis (B) abstracting protons from
the structure.

– e– +Nu– –e–
RH RH + RH(Nu) + RH(Nu) +Nu– Nu–RH–Nu

B
BH+ + R Radical chemistry

When the radical cation reacts with a nucleophilic species, the protonated radical
•RH(Nu) is generally more easily oxidized than the parent hydrocarbon and consequently
a cation + RH(Nu) is obtained, which by a second reaction with the nucleophile gives rise
to a disubstituted compound.
The reactions that can be involved in the oxidation of hydrocarbons are mainly those
of ET intermediates. In this framework, addition, substitutions, substitutions at α-carbons,
dimerization–elimination, and chain reactions are possible and they have been reviewed in
the literature [41].
k Another relevant group of compounds that can be oxidized are the carboxylates k
(RCOO− ), whose general reaction was first studied by Faraday in 1834 and Kolbe in 1849.
This reaction is the oldest and most studied process of the organic electrochemistry due
to its versatility, which consists of triggering between a radical (Kolbe reaction) and a
carbocationic chemistry (non-Kolbe reaction) [52]. This reaction has been mainly studied
under drastic conditions of electrolysis and nevertheless the mechanism is complex,
it is recognized that the first step afford a highly reactive acyloxy radical (RCOO• ),
which is further decomposed in carbon dioxide and a free radical (R• ). Depending on
the electrolysis conditions, this species can participate in a wide variety of dimerization
reactions, hydrogen elimination for the formation of alkanes and alkenes or reaction
additions to double bonds.

Carbocationic
chemistry

–e– –co2 –e–


RCO2– RCO2 R R+

Radical
chemistry

Due to the fact that the radicals are more easily oxidized than the original carboxylates,
carbocations can be formed at the oxidation potential of the carboxylate. In this way, a
wide variety of products have been obtained depending on the composition of the medium.

k
k

GENERAL ASPECTS OF REDOX CHEMISTRY 13

Thus, in the presence of water, alcohols or acetonitrile, alcohols, esters, or acetamides can
be synthesized, although the reaction between the carbocation and the carboxylate itself
can yield an ester.
The oxidation of compounds containing nitrogen can be done at the level of the unpaired
electrons to give rise to radical cations, which depending on the structure of the compound
can experiment different reactions. The simplest case of oxidation of amines concerns the
aliphatic compounds and even in this case, the mechanism and reaction products are diverse
and depend on the acid–base properties of the medium and the electrode material. High
acidity radical cations are formed after the first step of oxidation and usually, they present
hydrogen atom abstraction reactions [53].

R1–CH2–NR2

–e–

+ R2 = H + –H+
R1–CH2 + NH2 R1–CH2–NR2 R1–CH–NR2

–e–

+ H2O + + Nu–
R1–CHO + HNR2 R1–CH= NR2 R1–CH–NR2
–H+

–2e– Nu
Basic medium
–3H+ R2 = H

k R1–C ≡ N k

Primary amines (R1 –CH2 –NR2 ) are oxidized to a radical cation in the first step to yield
a relatively stable carbocation (R1 –CH2 + ) that can be decomposed by nucleophilic attack.
Alternatively, by hydrogen atom abstraction, the radical cation is transformed into a free
radical that through a second oxidation gives an iminium cation (R1 –CH = + NR2 ). In the
presence of water, the iminium cation can also be deprotonated (when R2 = H) to produce
aldehydes (R1 –CHO). By subsequent oxidation and deprotonation, the iminium cation can
finally transform into a nitrile structure (R1 CN). On the other hand, the oxidation of aro-
matic amines is complex and can produce a wide variety of products depending on their
structure and the use of either aqueous or nonaqueous medium or acid/basic conditions [54].
In general, other sorts of organic compounds containing unpaired electrons are also sus-
ceptible to be oxidized. For example, oxygen, sulfur as well as heterocyclic compounds,
which present a rich chemistry from the point of view of the interconversion of functional
groups and rearrangements [41]. All the organic reactions described in this section are acti-
vated by heterogeneous ET on an electrode, and in principle, they could also be done in
homogeneous phase by choosing middle and strong oxidants.

4 MODULATION OF REDOX POTENTIALS BY SUPRAMOLECULAR


INTERACTIONS

The success of ET processes between a substrate and an oxidant or reductant molecule


depends on different factors: the rate of homogeneous electron exchange in solution; the

k
k

14 ENCYCLOPEDIA OF PHYSICAL ORGANIC CHEMISTRY

driving force given by the exergonicity of the coupled chemical reactions following the for-
mation of reactive intermediates derived from the substrate; and the separation between the
redox potentials of the oxidant or reductant species and the substrate. Regarding the latter
factor, the values of these potentials depend on the electronic properties of the molecules
that exchange electrons; however, the potential modulation can be alternatively reached by
supramolecular interactions such as hydrogen bonding [55–58]. Other forces that can par-
ticipate in this kind of association processes are ion–ion-type electrostatic attractions (ion
pairing) and dipole–dipole associations provoked by interactions π–π between aromatic
rings, π-cation interactions [18, 55, 56, 58–61], van der Waals forces, and hydrophobic
effects. In the case of hydrogen-bonding interactions, the inter- or intramolecular modes can
be recognized and, respectively, induced by an appropriate control of the reaction medium
or structural modifications. The principle of this modulation is based on the supramolecu-
lar stabilization achieved by the ET product of the oxidant or reductant species, and this
possibility has been mainly analyzed from the point of view of electrochemical meth-
ods. In some cases, it comprised the use of sole or coupled techniques such as Raman
spectroscopy-voltammetry, ESR-voltammetry, and UV–vis voltammetry [62–65].
In this matter, quinone and nitroaromatic systems are a particular class of organic com-
pounds for which the redox potential modulation can be easily exemplified because its
reduction products can be stabilized by hydrogen-bonding interactions with weak proton
donors. When these compounds do not contain acidic groups in their structure, the reduc-
tion products are typically stable anions (radical anions and/or dianions) that interact with
the proton donor following one or more reversible association steps [58, 64, 66, 67].
The next set of reactions illustrates the association processes that can occur after the
one-electron reduction step of some oxidant species (O) in the presence of weak proton
k donors (DH): k
E

O + e −
↽−−⇀
−−−− −− O⋅− (11)
K1
O⋅− + DH −−−−−⇀

− −− O⋅− (DH) (12)
K2
O⋅− (DH) + DH −−−−−⇀

− −− O⋅− (DH2 )2 (13)

⋮ ⋮
Kn
O⋅− (DH)n−1 + DH −−−−−⇀

− −− O⋅− (DH)n (14)

——————————————————
Kglobal
O + e− + nDH ↽ ⇀
−−−−−−−−−− O⋅− (DH)n (15)
Even though several association steps can be involved, Nernstian models can be used
to explain the effect of the concentration of the additive DH on the redox potential of
the oxidant [58, 67], in most of the studied cases, a quinone or a nitrobenzene deriva-
tive. Equation 16 is the simplest model that describes the association as the global process
depicted by Equation 15, whereas Equation 17 is more general and considers that the dif-
ferent association equilibriums occur step by step (Equations 11–14).

RT
E1∕2 = E∘ + ln(1 + Kglobal [DH]n ) (16)
F

k
k

GENERAL ASPECTS OF REDOX CHEMISTRY 15

RT
E1∕2 = E∘ + ln(1 + K1 [DH] + K1 K2 [DH]2 + · · · + K1 K2 · · · Kn [DH]n ) (17)
F

These equations allow determining the equilibrium properties of the process and the
stoichiometry of the association and also show the combined effect of the values of the
equilibrium parameters and the concentration of the proton donor. According to this, the
half-wave potential (E1/2 ) of the overall process increases with the magnitude of the asso-
ciation and the proton donor concentration [67, 68]. In this way, if it were considered that
O is an oxidant for some particular substrate in a homogeneous redox process, the driving
force could be modulated by association by selecting appropriate conditions for the estab-
lishment of associative interactions. An equivalent situation can be reached in the case of
molecules in which the anions can also be stabilized by intramolecular interactions. Sev-
eral examples of this have been discussed in literature for the case of hydroxyl substituted
naphthoquinones [67, 69] and nitroaromatic compounds [70].
The modulation of the redox potential by supramolecular interactions is not limited for
quinones and nitroaromatic compounds as oxidants and hydrogen-bonding interactions, and
the concept can be used in a more general way for oxidants and reductants interacting in
all the modes mentioned previously [71–73].
The degree of shift in the redox potential of the oxidant is determined by the strength
of the hydrogen-bonding interaction; however, in the limit, formal proton transfer could
also occur. In this case, covalent bond formation can also contribute favorably to the total
free energy of the process making the initial redox process possible. For example, when
benzoquinone is used as oxidant in an acetonitrile solution containing solid copper, no
reaction occurs due to the great separation between the reduction potential of benzoquinone
k (<0.20 V/SCE) and the oxidation potential of copper (>0.20 V/SCE) [74, 75]; however, by k
adding a moderately strong proton donor like acetic acid, the respective hydroquinone and
the copper salt are formed. This example illustrates that in a redox process, the coupled
chemical reactions related to the oxidized form of substrate as well as the coupled reac-
tions of the reduced form of the oxidant can be determining factors of the feasibility of the
process. The examples here considered were for the case of an oxidant, hydrogen-bonding
interactions and proton transfer reactions, however, extrapolation to the case of reductants
and other supramolecular and chemical processes is straightforward [59, 60, 76–81].

REFERENCES

1. Taube, H., Myers, H., and Rich, R.L. (1953) Observations on the mechanism of electron transfer
in solution. J. Am. Chem. Soc., 75, 4118–4119.
2. Astruc, D. (1995) Electron Transfer and Radical Processes in Transition-Metal Chemistry,
Wiley-VCH, New York.
3. Marcus, R. (1997) Electron transfer reactions in chemistry: theory and experiment. J. Electroanal.
Chem., 438, 251–259.
4. Creutz, C. and Taube, H. (1969) Direct approach to measuring the Franck–Condon barrier to
electron transfer between metal ions. J. Am. Chem. Soc., 91, 3988–3989.
5. Costentin, C., Robert, M., and Savéant, J.M. (2006) Electron transfer and bond breaking: recent
advances. Chem. Phys., 324, 40–56.
6. Connelly, N.G. and Geiger, W.E. (1996) Chemical redox agents for organometallic chemistry.
Chem. Rev., 96 (2), 877–910.

k
k

16 ENCYCLOPEDIA OF PHYSICAL ORGANIC CHEMISTRY

7. Bard, A.J. and Faulkner, L.R. (2001) Electrochemical Methods, Fundamentals and Applications,
Wiley, New York.
8. Perrin, C.L. (1977) Necessity of electron transfer and a radical pair in the nitration of reactive
aromatics. J. Am. Chem. Soc., 99, 5516–5518.
9. Kornblum, N., Erickson, A.S., Kelly, W.J., and Henggeler, B. (1982) Conversion of nitro paraffins
into aldehydes and ketones. J. Org. Chem., 47, 4534–4538.
10. Ashby, E.C. (1980) A detailed description of the mechanism of reaction of Grignard reagents
with ketones. Pure Appl. Chem., 52, 545–569.
11. Thirsk, H.R. and Schmidt, P.P. (1975) Electron-Transfer Reactions in Specialist Periodical
Reports, Electrochemistry, vol. 5, Royal Chemical Society, London, pp. 21–131.
12. Thirsk, H.R. and Schmidt, P.P. (1978) The theory of electron-transfer reactions in polar media:
part II, in Specialist Periodical Reports, Electrochemistry, vol. 6, Royal Chemical Society, Lon-
don, pp. 128–242.
13. Smith, M.B. (2013) March’s Advanced Organic Chemistry, Reactions, Mechanisms and Struc-
ture, 7th edn, John Wiley & Sons, NJ, USA.
14. Wiberg, K.B. (1963) Oxidation–reduction mechanisms in organic chemistry. Surv. Progr. Chem.,
1, 211–248.
15. Litter, J.S. (1970) The mechanisms of oxidation of organic compounds with one-equivalent
metal-ion oxidants: bonded and nonbonded electron transfer, in Essays on Free Radical Chem-
istry (ed J.S. Littler), Royal Chemical Society, London, p. 383, Special Publication No. 24.
16. Litter, J.S. (1973) in Free Radical Reactions, International Review of Science, Organic Chemistry
Series One, vol. 10 (ed W.A. Waters), Butterworths, London, Chap. 8, p. 237.
17. Zuman, P. (1967) Substituent Effects in Organic Polarography, 1st edn, Plenum Press, USA.
18. Aguilar-Martínez, M., Bautista-Martínez, J.A., Macías-Ruvalcaba, N. et al. (2001) Molecular
structure of substituted phenylamine α-OMe- and α-OH-p-benzoquinone derivatives. synthesis
k and correlation of spectroscopic, electrochemical, and theoretical parameters. J. Org. Chem., 66,
k
8349–8363.
19. Engman, L., Persson, J., Andersson, C.M., and Berglund, M. (1992) Application of the Hammett
equation to the electrochemical oxidation of diaryl chalcogenides and aryl methyl chalcogenides.
J. Chem. Soc., Perkin Trans. 2, (8), 1309–1313.
20. Archer, J.F. and Grimshaw, J. (1969) Electrochemical reactions. Part VI. Application of the
Hammett relationship to the polarographic reduction of substituted 9-benzylidenefluorenes and
3-phenylcoumarins. J. Chem. Soc. B, 266–270.
21. Yamago, S., Kokubo, K., Hara, O. et al. (2002) Electrochemistry of chalcogenoglycosides. ratio-
nal design of iterative glycosylation based on reactivity control of glycosyl donors and acceptors
by oxidation potentials. J. Org. Chem., 67 (24), 8584–8592.
22. Chiu, K.Y., Su, T.H., Huang, C.W. et al. (2005) N,N,N′ ,N′ -Tetraaryl-p-phenylenediamine deriva-
tives. J. Electroanal. Chem., 578, 283–287.
23. D’Andrade, B.W., Datta, S., Forrest, S.R. et al. (2005) Relationship between the ionization and
oxidation potentials of molecular organic semiconductors. Org. Electron., 6 (1), 11–20.
24. Speelman, A.L. and Gillmore, J.G. (2008) Efficient computational methods for accurately pre-
dicting reduction potentials of organic molecules. J. Phys. Chem. A, 112 (25), 5684–5690.
25. Davis, A.P. and Fry, A.J. (2010) Experimental and computed absolute redox potentials of poly-
cyclic aromatic hydrocarbons are highly linearly correlated over a wide range of structures and
potentials. J. Phys. Chem. A, 114 (46), 12299–12304.
26. Sasaki, K., Kashimura, T., Ohura, M. et al. (1990) Solvent effect in the electrochemical reduction
of P-quinones in several aprotic solvents. J. Electrochem. Soc., 137, 2437–2443.
27. Sereda, G., Heukelom, J.V., Koppang, M. et al. (2006) Effect of transannular interaction on the
redox-potentials in a series of bicyclic quinones. Beilstein J. Org. Chem., 2 (26, http://www.
beilstein-journals.org/bjoc/content/pdf/1860-5397-2-26.pdf).

k
k

GENERAL ASPECTS OF REDOX CHEMISTRY 17

28. Marcus, R.A. (1956) On the theory of oxidation-reduction reactions involving electron transfer.
J. Chem. Phys., 24, 966–978.
29. Micaroni, L., Nart, F., and Hümmelgen, I. (2002) Considerations about the electrochemical esti-
mation of the ionization potential of conducting polymers. J. Solid State Electrochem., 7 (1),
55–59.
30. Frontana, C., Vázquez-Mayagoitia, A., Garza, J. et al. (2006) Substituent effect on a family of
quinones in aprotic solvents: an experimental and theoretical approach. J. Phys. Chem. A, 110
(30), 9411–9419.
31. Li, K., Li, M., and Xue, D. (2012) Solution reaction design: electroaccepting and electrodonating
powers of ions in solution. Nanoscale Res. Lett., 7, 6, http://www.nanoscalereslett.com/content/
7/1/6 (accessed March 2014).
32. Gritzner, G. (1991) Solvent effects on the redox properties of inorganic compounds. Rev. Inorg.
Chem., 11, 81–122.
33. Ransdell, R.A. and Wamser, C.C. (1992) Solvent and substituent effects on the redox prop-
erties of free-base tetraphenylporphyrins in DMSO and aqueous DMSO. J. Phys. Chem., 96,
10572–10575.
34. Tolbert, L.M. and Merrick, R.D. (1982) Carbanion photochemistry. 7. Electron transfer quench-
ing of carbanion photoreactivity by condensed aromatic hydrocarbons. J. Org. Chem., 47 (14),
2808–2810.
35. Arnett, E.M., Molter, K.E., Marchot, E.C. et al. (1987) Evidence for the equilibration of
resonance-stabilized carbocations, carbanions, and radicals by single electron transfer in solu-
tion. J. Am. Chem. Soc., 109, 3788–3789.
36. Mohammad, M., Khan, A.Y., Iqbal, M. et al. (1978) Pyridinyl radicals, pyridinyl anions, Z values,
and disproportionation equilibriums. J. Am. Chem. Soc., 100, 7658–7660.
37. Marcus, R.A. (1957) On the theory of oxidation-reduction reactions involving electron transfer.
k III. Applications to data on the rates of organic redox reactions. J. Chem. Phys., 26, 872–877. k
38. Nelsen, S.F., Hollinsed, W.C., Kessel, C.R., and Calabrese, J.C. (1978) Geometry change
upon electron removal from a tetraalkylhydrazine. X-ray crystallographic structures of
9,9′ -bis-9-azabicyclo[3.3.1]nonane and its radical cation hexafluorophosphate. J. Am. Chem.
Soc., 100, 7876–7882.
39. Kojima, H., Bard, A.J., Wong, H.N.C., and Sondheimer, F. (1976) Electrochemical
reduction of sym-dibenzocyclooctatetraene, sym-dibenzo-1,5-cyclooctadiene-3,7-diyne, and
sym-dibenzo-1,3,5-cyclooctatrien-7-yne. J. Am. Chem. Soc., 98 (18), 5560–5565.
40. Rehm, D. and Weller, A. (1970) Kinetics of fluorescence quenching by electron and H-atom
transfer. Israel J. Chem., 8, 259–271.
41. Lund, H. and Hammerich, O. (eds) (2001) Organic Electrochemistry, Dekker, New York.
42. Amatore, C., Gareil, M., and Savéant, J.M. (1983) Homogeneous vs heterogeneous electron
transfer in electrochemical reactions: application to the electrohydrogenation of anthracene and
related reactions. J. Electroanal. Chem., 147, 1–38.
43. Wawzonek, S. and Wearring, D. (1959) Polarographic studies in acetonitrile and dimethylfor-
mamide. IV. Stability of anion-free radicals. J. Am. Chem. Soc., 81, 2067–2069.
44. Lund, H., Michel, M.A., and Simonet, J. (1974) Homogeneous electron exchange in catalytic
polarographic reduction. Acta Chem. Scand., B28, 900–904.
45. Hawley, M.D. (1980) Encyclopedia of Electrochemistry of the Elements, Dekker, New York, pp.
1–308.
46. Steinberger, N. and Fraenkel, G.K. (1964) Electron spin resonance of the acetophenone and ben-
zaldehyde anion radicals. J. Chem. Phys., 40, 723–729.
47. Barnes, D. and Zuman, P. (1973) Polarographic reduction of aldehydes and ketones: XV. Hydra-
tion and acid-base equilibria accompanying reduction of aliphatic aldehydes. J. Electroanal.
Chem., 46, 323–342.

k
k

18 ENCYCLOPEDIA OF PHYSICAL ORGANIC CHEMISTRY

48. Andrieux, C.P., Grzeszczuk, M., and Savéant, J.M. (1991) Electrochemical generation and reduc-
tion of organic free radicals. α-Hydroxybenzyl radicals from the reduction of benzaldehyde. J.
Am. Chem. Soc., 113, 8811–8817.
49. Sugino, K. and Nonaka, T. (1965) Cathodic crossed hydrocoupling: II.
γ-Hydroxy-γ-methylvaleronitrile and its hydrolysis product from a mixture of acetone and
acrylonitrile. J. Electrochem. Soc., 112, 1241–1242.
50. Ikeda, Y. and Manda, E. (1985) Syntheses of benzilic acids through electrochemical reductive
carboxylation of benzophenones in the presence of carbon dioxide. Bull. Chem. Soc. Jpn., 58,
1723–1726.
51. Aguilar-Martínez, M., Macías-Ruvalcaba, N.A., Bautista-Martínez, J.A. et al. (2004) Hydro-
gen bond and protonation as modifying factors of the quinone reactivity. Curr. Org. Chem., 8,
1721–1738.
52. Schäfer, H.J. (1990) Recent contributions of Kolbe electrolysis to organic synthesis. Top. Curr.
Chem., 152, 91–151.
53. Barnes, K. and Mann, C.K. (1967) Electrochemical oxidation of primary aliphatic amines. J. Org.
Chem., 32, 1474–1479.
54. Weinberg, N.L. and Weinberg, H.R. (1968) Electrochemical oxidation of organic compounds.
Chem. Rev., 68, 449–523.
55. Gokel, G.W. (ed) (1990) Advances in Supramolecular Chemistry, vol. 1, JAI Press, USA, p. 197.
56. Jeffrey, G.A. and Saenger, W. (1991) Hydrogen Bonding in Biological Structures, Springer,
Berlin Heidelberg, Germany.
57. Ge, Y., Miller, L., Ouimet, T., and Smith, D.K. (2000) Electrochemically controlled hydrogen
bonding. o-Quinones as simple redox-dependent receptors for arylureas. J. Org. Chem., 65 (26),
8831–8838.
58. Gómez, M., González, F.J., and González, I. (2003) A model for characterization of successive
k hydrogen bonding interactions with electrochemically generated charged species. The quinone k
electroreduction in the presence of donor protons. Electroanalysis, 15 (7), 635–645.
59. Miyoshi, K., Oyama, M., and Okazaki, S. (2001) Electrochemical analysis of ion-pair formation
reactions involving organic dianions using differential pulse voltammetry. Electroanalysis, 13
(11), 917–922.
60. Brooksby, P.A., Schiel, D.R., and Abell, A.D. (2008) Electrochemistry of catechol terminated
monolayers with Cu(II),Ni(II) and Fe(III) cations: a model for the marine adhesive interface.
Langmuir, 24, 9074–9081.
61. Lebedev, A.V., Ivanova, M.V., and Ruuge, E.K. (2003) How do calcium ions induce free
radical oxidation of hydroxy-1,4-naphthoquinone? Ca2+ stabilizes the naphthosemiquinone
anion-radical of Echinochrome A. Arch. Biochem. Biophys., 413 (2), 191–198.
62. Yamanuki, M., Hocino, T., Oyama, M., and Okazaki, S. (1998) Thin layer electrochemical Raman
study of ion pair formation between the tetrachlorobenzoquinone anion radical and alkaline earth
metal cations. J. Electroanal. Chem., 458, 191–198.
63. Batanen, V., Eloranta, J.M., and Vuolle, M. (1998) EPR and ENDOR spectroscopy studies on
α-aminoanthraquinone radical cations in solution. J. Chem. Soc., Perkin Trans. 2, 2483–2488.
64. Gupta, N. and Linschitz, H. (1997) Hydrogen-bonding and protonation effects in electrochem-
istry of quinones in aprotic solvents. J. Am. Chem. Soc., 119, 6384–6391.
65. Tang, Y., Wu, Y., and Wang, Z. (2001) Spectroelectrochemistry for electroreduction of
p-benzoquinone in unbuffered aqueous solution. J. Electrochem. Soc., 148, E133–E138.
66. Uno, B., Kawabata, A., and Kano, K. (1992) How active p-quinone dianions are as proton accep-
tors!. Chem. Lett., 21 (6), 1017–1020.
67. Gómez, M., González, F.J., and González, I. (2003) Effect of the host and guest structures on
hydrogen bonding association. influence on the stoichiometry and equilibrium constants. J. Elec-
trochem. Soc., 150 (11), E527–E534.

k
k

GENERAL ASPECTS OF REDOX CHEMISTRY 19

68. Galano, A., Gómez, M., González, F.J., and González, I. (2012) Correlation between hydrogen
bonding association constants in solution with quantum chemistry indexes. The case of successive
association between reduced species of quinones and methanol. J. Phys. Chem. A, 116 (43),
10638–10645.
69. Macías-Ruvalcaba, N.A., González, I., and Aguilar-Martínez, M. (2004) Evolution from hydro-
gen bond to proton transfer pathways in the electroreduction of α-NH-quinones in acetonitrile.
J. Electrochem. Soc., 151 (3), E110–E118.
70. Truong, D., Woods, J.E., Bu, J. et al. (2005) Electrochemically controlled hydrogen bonding.
Nitrobenzenes as simple redox-dependent receptors for arylureas. J. Am. Chem. Soc., 127 (17),
6423–6429.
71. Jordan, B.J., Pollier, M.A., Miller, L.A. et al. (2007) Redox-modulated recognition of tetrazines
using thioureas. Org. Lett., 9 (15), 2835–2838.
72. Salas, M., Gómez, M., González, F.J., and Gordillo, B. (2003) Electrochemical reduction of
1,4-benzoquinone. Interaction with alkylated thymine and adenine nucleobases. J. Electroanal.
Chem., 543 (1), 73–81.
73. Gómez, M., Gómez-Castro, C.Z., Padilla-Martínez, I.I. et al. (2004) Hydrogen bonding effects
on the association processes between chloranil and a series of amides. J. Electroanal. Chem., 567
(2), 269–276.
74. Libuś, W. and Strzelecki, H. (1972) Potentiometric study on the activity coefficients in equimolal
mixtures of some divalent metal perchlorates in acetonitrile solution. Electrochim. Acta, 17 (3),
577–585.
75. MacLeod, I.D., Parker, A.J., and Singh, P. (1981) Electrochemistry of copper in aqueous ace-
tonitrile. J. Solution Chem., 10 (11), 757–774.
76. Armstrong, F.A., Hill, H.A.O., and Walton, N.J. (1982) Direct electrochemical reduction of ferre-
doxin promoted by Mg2+ . FEBS Lett., 145 (2), 241–244.
k k
77. Astudillo, P.D., Tiburcio, J., and González, F.J. (2007) The role of acids and bases on the elec-
trochemical oxidation of hydroquinone: hydrogen bonding interactions in acetonitrile. J. Elec-
troanal. Chem., 604 (1), 57–64.
78. Okamura, T., Iwamura, T., Yamamoto, H., and Ueyama, N. (2007) Synthesis and molecular struc-
tures of S-2-FcNHCOC6 H4 SH and [MIII(OEP)(S-2-FcNHCOC6 H4 )] (Fc = ferrocenyl, M = Fe,
Ga): electrochemical contributions of intramolecular SH· · ·OC and NH· · ·S hydrogen bonds. J.
Organomet. Chem., 692 (1–3), 248–256.
79. Alligrant, T.M., Hackett, J.C., and Alvarez, J.C. (2010) Acid/base and hydrogen bonding effects
on the proton-coupled electron transfer of quinones and hydroquinones in acetonitrile: mech-
anistic investigation by voltammetry, 1 H NMR and computation. Electrochim. Acta, 55 (22),
6507–6516.
80. Noël, V. and Randriamahazaka, H.N. (2012) Redox-assisted hydrogen bonding within interpen-
etrating conducting polymer networks for charge-storage materials. Electrochem. Commun., 19,
32–35.
81. Peters, A., Rainka, M.P., Krishnan, L. et al. (2013) Electrochemical characterization of
hydrogen-bonding complexation between indoline and nitrogen containing bases. J. Electroanal.
Chem., 691 (15), 57–65.

Potrebbero piacerti anche