You are on page 1of 16

Coordination Chemistry Reviews 309 (2016) 68–83

Contents lists available at ScienceDirect

Coordination Chemistry Reviews


journal homepage: www.elsevier.com/locate/ccr

Review

Redox chemistry of cobalamin and its derivatives


I.A. Dereven’kov a,∗ , D.S. Salnikov a , R. Silaghi-Dumitrescu b , S.V. Makarov a , O.I. Koifman a
a
Institute of Macroheterocyclic Compounds, Ivanovo State University of Chemistry and Technology, Sheremetevskiy str. 7,
Ivanovo 153000 Russian Federation
b
Department of Chemistry, Babes-Bolyai University, Arany Janos 11, Cluj-Napoca RO-400024 Romania

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
2. Dependence of the coordination properties of cobalamins and their derivatives on the oxidation state of the cobalt ion . . . . . . . . . . . . . . . . . . . . . . . . . . 69
3. Mechanisms for the reduction reactions of the Co(III) forms of cobalamins and their derivatives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70
4. Mechanisms of reduction reactions for Co(II) forms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
5. Mechanisms of oxidation of the Co(II) forms of cobalamins and their derivatives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
6. Oxidation mechanisms for Co(I) cobalamins and their derivatives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
7. Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81

a r t i c l e i n f o a b s t r a c t

Article history: A comprehensive review on redox reactions of Co(III), Co(II) and Co(I) forms of cobalamins (vitamin B12)
Received 8 September 2015 and their derivatives (cobinamide, cobyrinates) is presented. An influence of coordination environment
Received in revised form 31 October 2015 of cobalt ion on reactivity of corresponding species is discussed. Mechanisms of reduction of Co(III) and
Accepted 1 November 2015
Co(II) species as well as oxidation of Co(II) and Co(I) complexes are given in terms of their in vivo and
Available online 14 November 2015
catalytic implications.
© 2015 Elsevier B.V. All rights reserved.
Keywords:
Cobalamin
Vitamin B12
Free radicals
Reduction
Oxidation
Reaction mechanisms

1. Introduction studies. Several general reviews are available [1–3], as well as work
summarizing interim results in relatively narrow areas: chemical
Cobalamins (Cbls; Fig. 1) are the most common in vivo cobalt modification of the structure of cobalamin [4] and their peptide
compounds, as well as the most complex vitamins in their struc- derivatives [5], Cbl-dependent enzymes [6], transport and pro-
ture – and have hence been the subject of various and extended cessing cobalamins in vivo [7–9], activation of Co C bonds in
adenosylcobalamin-dependent enzymes [10], catalytic reactions
involving Cbls [11], corrin-based antivitamins [12,13], medical
applications of cobalamin derivatives [14] and others. Neverthe-
Abbreviations: Cbl, cobalamin; Cbi, cobinamide; Ado, 5 -deoxyadenosyl;
less, there have been no reviews dedicated to the redox reactions
ATP, adenosine triphosphate; bpy, 2,2 -bipyridine; CASPT2, complete active
space perturbation theory; CASSCF, complete active space self-consistent field; of the cobalamins.
CoA, coenzyme A; DDT, 1,1-bis(4-chlorophenyl)-2,2,2-trichloroethane; DMBI, 5,6- Cobalamin transport in mammals [7–9], its ligation changes
dimethylbenzimidazole; DMF, dimethyl formamide; DMSO, dimethyl sulfoxide; [7–9], the synthesis of coenzyme forms [6,9] and, in fact, Cbl-
DFT, density functional theory; GSH, l-glutathione; HMS, hydroxymethanesulfinate; dependent enzymatic processes [6,10] all include as a main event
LMCT, ligand to metal charge transfer; LUMO, lowest unoccupied molecular orbital;
MOF, metal-organic frameworks; NHE, normal hydrogen electrode.
oxidation or reduction of the cobalt ion. Under physiological con-
∗ Corresponding author. Tel.: +7 4932 327357; fax: +7 4932 417995. ditions, Co(III), Co(II), and Co(I) forms of cobalamins are known.
E-mail address: derevenkov@gmail.com (I.A. Dereven’kov). Cbl(I) is the necessary form for the synthesis or organo-cobalamins

http://dx.doi.org/10.1016/j.ccr.2015.11.001
0010-8545/© 2015 Elsevier B.V. All rights reserved.
I.A. Dereven’kov et al. / Coordination Chemistry Reviews 309 (2016) 68–83 69

Fig. 1. Structures of cobalamin (A), cobinamide (B), cobyrinic acid (C) and heptamethyl cobyrinate (cobester) (D).

(Co(III)-complexes) involved in the transfer of a methyl moiety 2. Dependence of the coordination properties of
(methylcobalamin, MeCbl) and isomerization of various substrates cobalamins and their derivatives on the oxidation state of
(adenosylcobalamin, AdoCbl). Deactivation of Cbl(I) leads to the the cobalt ion
formation of Cbl(II), which is also formed by homolysis Co C bond
in AdoCbl-dependent enzymes. Cobalamin redox reactivity is strongly dependent on the coordi-
Catalytic conversion of various organic substrates is also pos- nation state of the cobalt ion. Coordination properties directly affect
sible using cobalamin derivatives (e.g., cobyrinates, cobesters, both the redox potential of the Co(III)/Co(II) and Co(II)/Co(I) pairs,
cobinamides (Cbi); Fig. 1) [11,15,16]. In such cases again, the key and the availability of coordination sites necessary for engaging
steps involve redox transitions between the Co(III), Co(II) and Co(I) into a complex with the reducing agent.
forms of these complexes. The mechanisms of these reactions are Cbls(III) (Fig. 1) are complexes where the corrin nitrogen atoms
quite diverse, but this is not sufficiently reflected in the currently occupy the four equatorial coordination sites, while the fifth coor-
available reviews. This article summarizes the redox chemistry dination site at the bottom (␣) axial position is either occupied by a
of cobalamins and their derivatives, including the reduction pro- nitrogen atom of the 5,6-dimethylbenzimidazole (DMBI) unit of the
cesses on Co(III) and Co(II) forms, as well as oxidation of Co(I) and nucleotide chain (in the base-on form of cobalamin) or by an oxy-
Co(II). gen atom of a water molecule (in the base-off form of cobalamin, or
70 I.A. Dereven’kov et al. / Coordination Chemistry Reviews 309 (2016) 68–83

Fig. 2. Coordination states of Co(III), Co(II) and Co(I)-corrins.


Source: Adapted from Ref. [3].

in incomplete corrinoids). The sixth coordination site (␤-position)


is available for coordination of various exogenous groups (X = H2 O,
CN− , CH3 − and others).
It is believed that Co(III) corrinoids are mostly hexacoordinated,
while the Co(II) form tends to be pentacoordinated, and Co(I) –
tetracoordinated (Fig. 2) [3].
It is important to mention that there are a number of exceptions
to this rule. For example, an axial ligand with a strong trans-
influence will lead to pentacoordinated Co(III)–corrin complexes.
Thus, the pressure dependence of UV–Vis spectra of MeCbl and Fig. 3. Hydrogen bonding of cobalamin(I).
vinyl–cobinamide (vinyl–Cbi(III)) (where the methyl and vinyl lig- Source: Adapted from Refs. [28,30,31].
ands have strong electron-donating properties) can be explained by
the equilibria hexacoordinated base-on MeCbl ↔ pentacoordinated Co(II)–Nax (DMBI) bond compared to the original Cbl(II)base-on .
base-off MeCbl, and hexacoordinated vinyl–Cbi(III)–H2 O ↔ pen- DFT calculations do not exclude the possibility of the formation of
tacoordinated vinyl–Cbi(III), which are displaced toward the hexacoordinated complexes of Cbl(II) with SCN− , where one of the
hexacoordinated state with increasing pressure [17]. Similarly, axial bonds is considerably weakened. Ligation of thiocyanate in
nitroxylcobalamin (Cbl(III)–NO− ) in neutral and alkaline aqueous this case was predicted by DFT data to occur more preferably via
solutions features significant amounts of the deprotonated base-off the nitrogen atom than via the sulfur atom, as also observed for
form, as demonstrated by 13 C and 31 P NMR [18]. Since the nitroxyl Co(III)-complexes of cobalamin and cobinamide [23].
anion exhibits a strong trans-influence [18], it can be assumed that There are also known examples of tetracoordinated Cbl(II).
the base-off form is indeed pentacoordinated. Spectroscopic and X-ray data on the binding of cobalamin(II) to
For Cbl(II) there are also a few exceptions known from the ATP:cobalamin adenosyltransferase, the enzyme that synthesizes
classic pentacoordinated state [19–21]. In the early stages of the AdoCbl in vivo, supports Co(II) state with very weak axial interac-
reaction between Cbl(II) and dithionite (S2 O4 2− ), a loss of the tions [24–26].
EPR signal of the starting Co(II) is observed; the resulting state Despite the fact that Cbl(I) is considered to be tetracoordinatied
is very stable in the absence of oxygen, and is oxidized by O2 to based on spectral data [21,27], the absence of crystallographic data
Cbl(III)–SO3 2− . The most adequate representation of this complex does not allow to consider this point of view unique. DFT calcula-
would be Co(II)–SO2 •− , where the unpaired electrons of Co(II) and tions predict a possibility of the formation of one or two hydrogen
SO2 •− are coupled to each other. This complex has a pKa = 4.8 at bonds between Co(I) and H-donor (H2 O [28–30], protonated imi-
25 ◦ C, corresponding to a base-on ↔ base-off transitions and imply- dazole [28], aminoacids such as tyrosine [31]; Fig. 3).
ing that at least under certain conditions Co(II) is hexacoordinated
Co(II) [22]. 3. Mechanisms for the reduction reactions of the Co(III)
A hexacoordinated complex of cobalamin(II) is also obtained forms of cobalamins and their derivatives
with SCN− . Thus, Cbl(II) is capable of binding thiocyanate in an
alkaline medium (K = 3.6 ± 0.3 M−1 at 25 ◦ C, I = 1 M). The UV–Vis Conversion of aqua-Co(III)-corrinoids to Co(II) complexes occurs
spectrum of the resulting complex is markedly different from that relatively easily. On one hand, this is due to the sufficiently high
of thiocyanato-cobinamide(II), which speaks of their structural potential of the Co(III)/Co(II)-couples. On the other hand, the high
differences (while, as expected, the UV–Vis spectra of the base-off mobility of water ligand facilitates its substitution by the reducing
complexes Cbl and Cbi are very similar to each other). Significant agent. The redox potentials of the H2 OCbl(III)base-on /Cbl(II)base-on
changes were also observed in the structure of Cbl(II) EPR signal and (H2 O)2 Cbi(III)/Cbi(II) pairs are +0.20 and +0.51 V at 22 ◦ C
upon addition of thiocyanate. The Cbl(II)–SCN− complex displays (relative to the normal hydrogen electrode, NHE) [32], which is
a typical acid–base equilibrium with pKa = 3.4 a 25 ◦ C, in which higher than those of the potential biological partners in reduction
only a protonation of bound DMBI can occur, since thiocyanate is reactions.
protonated in strongly acidic medium. These facts indicate the for- Redox potential of Cbl(III) is strongly dependent on pH: a value
mation of a hexacoordinated complex with a significantly labilized of +0.20 V (22◦ C, NHE), remains independent only in the region
I.A. Dereven’kov et al. / Coordination Chemistry Reviews 309 (2016) 68–83 71

Scheme 1. A common mechanism of reduction of aquacobalamin.


Source: Adapted from Ref. [22].

between pH 2.9 and 7.8 and moves to more negative values above monosaccharides possess reductive activity only at strongly alka-
pH 7.8 and more positive values below pH 2.9 [32]. Latter can be line pH – so that pH value of 12–13 are preferred for the actual
explained by acid-base transformations of axial ligand depending reduction of the cobalt, after which the reactivity of the product
on basicity of medium: at pH >7.8 aqua ligand is deprotonated to may be explored at ease using lower pH values, where the excess
hydroxide (reaction (1)), at pH <2.9 DMBI attaches proton giving monosaccharides are redox-inactive enough to avoid interferences.
Cbl(II) and Cbl(III) base-off-species (Fig. 2). However, the most convenient route to Co(II) as well as to Co(I)
Several suitable reducing agents for the aqua-forms of Cbl(III) is reduction with sodium borohydride [36]. Depending on the con-
and Cbi(III) to the corresponding Co(II)-complexes are known – centration of the reducing agent, the reaction may be stopped
including ascorbic acid [33], monosaccharides [34,35], sodium at Co(II) or continued on to Co(I). The excess reducing agent can
borohydride[36], sodium formate [37], and thiols [38]. be quenched either by performing the reduction in weakly acidic
Ascorbic acid is an effective reductant of aquacobalamin, with media (e.g., pH 4–5, where within about 1 h the remaining NaBH4
a rate of 24 M−1 s−1 at pH 7.0; 25 ◦ C [33]. However, the oxidation decomposes generating H2 and borate derivatives), or by adding
product of ascorbate, dehydroascorbic acid, is sufficiently stable acetone after completion of the desired reduction reaction.
in neutral and acidic solution and has oxidizing properties, which Aqua Cbl(III) and Cbi(III) react with thiols via rather complicated
may negatively influence the course of experiments on the reac- mechanisms, with various products depending on the nature of the
tivity of Co(II)-corrinoids. Under certain conditions, destruction of starting reactants. With Cbl(III)–OH2 , glutathione (GSH) [40,43], N-
the macrocycle was also reported [33,39]. Notably, this destruction acetylcysteine [44], homocysteine [44], ␥-glutamylcysteine [45],
refers primarily to experiments conducted under aerobic condi- or captopril [46] yield reasonably stable Cbl(III)–SR− complexes,
tions, where the inevitably produced hydrogen peroxide is, as most whose structure was established by X-ray diffraction. These thi-
reactive oxygen species, likely to destroy corrin [33]. olate complexes are stable in neutral, weakly acidic and weakly
Reduction of aquacobalamin by formate [37] depends on the alkaline solutions. In a strongly acidic medium a slow hydrolysis
acid–base properties of two reagents: there was a decrease in rate to the original RSH and Cbl(III)–OH2 is observed [47]. At neutral
at pH <5 in due to formate protonation (pKa = 3.53 [37]) and at pH >7 pH, the UV–Vis spectra of these complexes vary little even after
due to Cbl(III)–OH2 deprotonation to Cbl(III)–OH− (pKa = 7.75 [40]). several hours in the presence of a large excess of the corresponding
The hydroxide ligand is inert to substitution, so that the reduction thiol. In alkaline media, reduction yields Cbl(II) and the correspond-
of Cbl(III)–OH− requires preliminary protonation (Scheme 1). It is ing disulfide, and requires a minimum of two equivalents of thiol
also expected that the reduction of Cbl(III)–OH2 by formate involves [48]. EPR data suggest that it is possible to capture thiyl radical by
hydride transfer, yielding Cbl(I), which rapidly comproportionates Co(II)-complexes [48].
with Cbl(III)–OH2 by reaction (1) [41]. However, direct evidence Thiolatocobalamins exhibit so-called “atypical” UV–Vis spectra,
for the formation Cbl(I) system is not available. Formate has a similarly to organo-Cbls(III), sulfito-Cbl(III) [49], phenolato-Cbl(III)
significant drawback for synthesis of Cbl(II): in order to achieve [50], and some other species, which are distinct from “typical”
optimal rate of reduction, relatively high concentrations of formate spectra of Cbl(III)–OH2 , Cbl(III)–CN− and others. A major difference
are needed (>10 mM). between these groups is in position of ␥-band, which is typically
located at 350 nm, but red-shifted and weakly structured in “atyp-
Cbl(I) + Cbl(III)–OH2 → 2Cbl(II) + H2 O. (1)
ical” spectra. DFT computations performed on Cbl(III)–SG− models
Cbl(III)–OH2 is reducible with glucose in alkaline medium [34]. [51], [52] reveal similarity between Co(III) S and Co(III) C bonds,
The reaction proceeds at a low rate and yields Cbl(II). The active i.e., GS− and CH3 − groups strongly decrease effective charge on Co
form of the reducing agent is the glucose anion, only available ion via ␴-donation and substantially shifts energy of 3d-orbitals to
under strongly alkaline medium – where, however, the concentra- more positive values. High energy of 3d-orbitals enables their mix-
tion of the active form of cobalamin is very low, which explains the ing with corrin ␲-orbitals increasing number of electron transitions
very low rates of the process. Reduction of dihydroxycobinamide in ␥-region with similar energies and explains appearance of “atyp-
[35] under these conditions proceeds at a higher rate and com- ical” spectra [51]. In Cbl(III)–OH2 or Cbl(III)–CN− , a sharp ␥-band of
prises two successive stages: Co(III) → Co(II) and Co(II) → Co(I). The UV–Vis spectrum is generally contributed by ␲ → ␲* corrin-based
higher rate of the first stage is due to the greater availability of the transitions, since their energy exceeds that of Co 3d → ␲* corrin-
aquahydroxo form (pKa for transition to dihydroxo: 10.3 at 25 ◦ C). based excitations.
In a strongly alkaline media the mechanism of the first reduction Reduction of Cbl(III)–OH2 by l-cysteine occurs more easily than
step is also changed: while at pH 11 reduction by excess glucose or with the above thiols [38], and is possible even in a weakly acidic
fructose is described by an exponential equation, the shape of the medium. Likewise facile is reduction of Cbl(III)–OH2 by dithiotreitol
kinetic curve becomes typical for an autocatalytic reaction at pH 13. (DTT) [48].
Apparently, the one-electron oxidized versions of the monosaccha- The interaction of Cbl(III)–OH2 [53] and (H2 O)2 Cbi(III) [54] with
rides, which, probably, are more efficient reducing agents for Co(III), hydrogen sulfide is a special case, with distinctly more complex
are more stable in such strongly alkaline media and can accumu- mechanisms than in the case of the thiols discussed above. In
late to amounts which allow them to play an important part in the acidic and neutral media under anaerobic conditions, aquacobal-
reduction of the cobalt. amin reacts with hydrogen sulfide in three steps (Scheme 2) [53].
Reduction by monosaccharides is a convenient method of The first stage entails rapid substitution of water by hydrogen
obtaining reduced forms of cobalamin derivatives [42], since sulfide, which proceeds at a higher rate than the formation of
72 I.A. Dereven’kov et al. / Coordination Chemistry Reviews 309 (2016) 68–83

subsequent conversion to known complexes (for example, methyl


or dicyano derivatives). However, if this reaction is carried out in the
presence of oxygen, modified corrin complexes are observed [53].
Interestingly, interaction of diaquacobinamide with hydrogen
sulfide in alkaline medium in an equimolar ratio leads to the for-
mation of a relatively stable complex, the UV–Vis spectrum of
which resembles Cbi(II) with the notable addition of a relatively
intense absorption band at 778 nm, which is extremely rarely seen
in corrin systems [54]. It is assumed that the long-wavelength
absorption bands in the complexes correspond to corrinoid ligand-
to-metal charge transfers (LMCT). These absorption bands have
been observed in complexes of cobalamin with I− , Br− , Cl− and
Scheme 2. A mechanism of reaction between aquacobalamin and hydrogen sulfide.
thiourea [56]. A 634-nm band also arises in a cobyrinate complex
Source: Adapted from Ref. [53].
with pentafluorophenylthiolate. Probably the strong polarization
of the Co(III) S bond affects energy of Co 3dz2 orbital and decrease
thiolate complexes under the same conditions. For example, the energy of corrin ␲ → Co 3dz2 transition corresponding to red region
second-order rate constant with hydrogen sulfide at pH 4.5, 25 ◦ C is of UV–Vis spectrum [57]. Apparently, the Cbi(II)–S•− complex also
(66 ± 1) M−1 s−1 , as opposed to (18.1 ± 0.1) M−1 s−1 for glutathione displays notable charge transfer between the metal and the axial
(at pH 5.0, 25 ◦ C). The reaction of aquacobalamin with hydrogen sul- ligand.
fide entails both molecular H2 S and the deprotonated form, HS− – Understanding the interaction of Cbl(III) and Cbi(III) with H2 S is
with rate constants of (58 ± 3) and (527 ± 89) M−1 s−1 , respectively, also relevant for the use of these metal complexes as antidotes of
which is consistent with the higher nucleophilicity of the HS− form. hydrogen sulfide. Hydrogen sulfide is a potent toxin originated from
For glutathione, the rates of reaction with three forms – protonated a number of natural and industrial sources inhibiting cytochrome
at the thiol, deprotonated at the thiol and protonated at the amino c oxidase in human body. Several recent studies report that high
group, and deprotonated at the thiol as well as at the amino group, doses of Cbl(III)–OH2 can decrease a mortality associated with
are 18.5, (28 ± 10) and (163 ± 8) M−1 s−1 , respectively [40]. Com- H2 S exposure [58–60]. Since oxidation of H2 S in vivo occurs in
parison of the rate constants for the individual steps shows higher time-scale of several minutes, administration of aquacobalamin
reactivity for H2 S and HS− compared to GSH and GS− . in poisoned victims should be carried out immediately (1–4 min
During the next two steps of the cobalamin-sulfide the changes after contact with H2 S) [60]. Furthermore, some forms of Cbi(III)
in the UV–Vis spectrum are suggestive of cobalt reduction, but (viz., aquahydroxo-, sulfito- and dinitro-species of Cbi(III)) are also
the spectrum of the final product is clearly different from Cbl(II). effective antidotes for hydrogen sulfide with efficacy exceeding
It is suggested that during these stages electron transfer occurs that of aquacobalamin [61]. Moreover, Cbi(III)-species are more
from hydrogen sulfide to Co(III), with the resulting sulfide radi- advantageous than Cbl(III)–OH2 due to the higher water solubility
cal remaining within the inner coordination sphere of the resulting of the former [61]. Probable mechanisms of action of Cbl(III) and
Co(II). Since the sulfide radical reacts at high rates with the remain- Cbi(III) complexes as antidotes includes binding of H2 S to Co(III)
ing hydrogen sulfide (reactions (2) and (3)) [55], it is assumed that [62], oxidation of H2 S by Co(III) [63] and others [60]. According
during the last stage of the Cbl–sulfide interaction a reversible to our data, both coordination and oxidation pathways occur in
interaction of the bound sulfide radical bound with a second H2 S/Co(III)-reaction, i.e., Co(III)-complexes readily bind one [53]
molecule of hydrogen sulfide occurs, which can lead to a shift in or two H2 S molecules [54] followed by slower reduction of Co(III)
the equilibrium toward pentacoordinated Cbl(II). to Co(II) [53,54]. Nevertheless, these studies [53,54] were per-
formed under anaerobic conditions and an influence of oxygen on
HS− + S•− ↔ HSS•2− . (2) H2 S/Co(III)-system remain unassessed. On the other hand, refer-
HS −
+ HS• ↔ HSSH•− . (3) ence [64] questions the detoxifying properties of the corrinoids
due to binding of free hydrogen sulfide, since most of the hydro-
Interaction of diaqua-cobinamide with hydrogen sulfide [54] gen sulfide within the blood stream would rapidly be subjected
occurs much faster than the reaction involving Cbl(III)–OH2 . More- to oxidation by plasma proteins [63]. Probably, further considera-
over, both axial water molecules may be replaced by hydrogen tion of the interaction between corrinoids and persulfides and thiyl
sulfide. Substitution of the first H2 O occurs at a high rate constant radicals would allow elucidation of these mechanisms of action of
– at pH 4.5, 25 ◦ C it is (2112 ± 171) M−1 s−1 , and at pH 9.6, 25 ◦ C it is cobalamin and cobinamide.
(11,051 ± 615) M−1 s−1 , which is more than an order of magnitude Complexes of cobalamin [65] and heptamethylcobyrinate [57]
greater than for the corresponding reactions of Cbl(III)–OH2 . Inter- with pentafluoriphenylthiolate display stabilities intermediate
action of the monohydrosulfide complex with a second molecule between those of complexes with biological thiols and hydrogen
of H2 S occurs at a lower rate (rate constant 81.7 M−1 s−1 at pH 9.5, sulfide, respectively. The cobalamin–pentafluorophenylthiolate
25 ◦ C), which may be explained by the fact that at this pH a signifi- complex is stable at −15 ◦ C in methanol; at room temperature,
cant fraction of the complex is in the (HS− )(HO− )Cbi(III) form, inert however, Cbl(II) and the corresponding disulfide are formed [65].
to substitution. Synthesis of cobirinate complexes with pentafluorophenylthiolate
In two subsequent stages electron transfer occurs from sulfide at room temperature is possible, but photolysis or thermolysis of
ion to Co(III), resulting in the formation Cbi(II)–S•− , and then react- the complex is also observed [57].
ing the latter with a second molecule of hydrogen sulfide, as in the The reactions of aquacobalamin with the stronger sulfur-
case of Cbl(III)–OH2 . Electron transfer is more facile with complexes containing reducing agents dithionite (S2 O4 2− ) [22], sulfoxylate
containing one molecule of hydrogen sulfide. Probably, addition of (SO2 2− ) [66], and hydroxymethanesulfinate (HOCH2 SO2 − ) have
the second H2 S molecule to the complex increases the nucleophilic- also been described [67]. As noted above, the reaction between
ity of the system, which leads to more negative redox potentials. Cbl(III)–OH2 and S2 O4 2− leads to the hexacoordinated Cbl(II)
It is important to underline that the reaction of hydrogen sul- complex with the radical SO2 •− . This complex is also formed
fide with cobalamin and cobinamide under anaerobic conditions during the reaction Cbl(III)–OH2 with sulfoxylate [66] (SO2 H− ,
does not lead to modifications of the macrocycle, as indicated by obtained by decomposition of thiourea dioxide in an alkaline
I.A. Dereven’kov et al. / Coordination Chemistry Reviews 309 (2016) 68–83 73

medium under anaerobic conditions [68]), as well as with hydrox- bond, leading to the formation Cbl(I) species which is not directly
ymethanesulfinate [67]. The final product in the reduction with detected in the system, arguably due to its high reactivity – so high
hydroxymethanesulfinate in strongly alkaline medium is Cbl(I). that even protons may oxidize it to Cbl(II).
With dithionite, the reducing agent is the SO2 •− radical, formed In the case of base-off complexes, ESI–MS data have confirmed
in reaction (4) from S2 O4 2− [69]. the formation of the six-coordinate thiolate complex preceding
deligation [83]. However, it is unlikely that the incoming trans-
S2 O4 2− ↔ 2SO2 •− . (4)
thiolate has a significant influence on the activation of the Co C
Reduction of cobalamin derivatives containing strongly bound bond, since the strength of the latter is practically independent
ligands at the ␤-position (e.g., CN− , CH3 − , Ado− and others), on the presence of ligands in the ␣-position [84,85]. Pyridine and
is much slower than for the aquacomplexes. Electrochemical Zn2+ do affect the reaction rates, presumably by facilitating thiol
reduction of cyanocobalamin is observed at −0.76 V (vs. NHE, deprotonation [83].
in DMSO/propanol), which corresponds to an irreversible two- Systems such as the CblC-trafficking protein illustrate deligation
electron process, i.e., the reaction product is Cbl(I) [70]. Reduction reactions of alkylcobalamins in biological systems. In this case, a
of organocobalamins occurs at even lower values of potentials. For transition to the base-off form occurs upon binding to the protein
example, the irreversible reduction of MeCbl occurs at −1.20 V (vs. [86,87]. This particular reaction requires GSH, which is eventually
NHE, 20 ◦ C), while AdoCbl is reduced at −1.04 V (in DMF/methanol) converted into the corresponding thioether. It is assumed that the
[71]. DFT calculations have correlated such redox potential obser- mechanism involves a nucleophilic substitution of the alkyl moiety
vations for alkylcobalamins with the energy of the LUMO [72]. by the thiolate, which leads to the formation of Cbl(I). While the
The kinetics of the reaction between Cbl(III)–CN− and dithionite CblC catalytic processes is more than 10,000 times faster than the
or hydroxymethanesulfinate were also analyzed [73]. With S2 O4 2− uncatalyzed one, the reason for such a significant catalytic effect is
the reaction rate is pH-independent, while in the case of hydrox- not clear [87].
ymethanesulfinate the rate increases with the pH. Concentration An interesting result was obtained in the study of the inter-
dependences have a non-linear shape and reach a plateau at high action between AdoCbl bound to the Caenorhabditis elegans CblC
concentrations of the reducing agent. Interestingly, the rate con- protein, and GSH [88]: Cbl(I) is formed at a relatively high rate, in
stants at the plateau phase for S2 O4 2− and HOCH2 SO2 − are identical contrast with the fact that free AdoCbl reacts only very slowly with
to each other (0.03 s−1 at 25 ◦ C). It is important to emphasize GSH. Moreover, excess glutathione significantly contributed to the
that this value is close to the dissociation rate constant for DMBI stabilization of the Cbl(I) form in the presence of oxygen. Proba-
Cbl(III)–CN− (0.042 s−1 at 25 ◦ C) [74]. Additionally, the activation bly, further structural studies of the CblC/Cbl complex may help
parameters are also identical for the reduction reaction and for understand the cause of this efficient activation of cobalamin by an
DMBI elimination. This indicates that the rate-determining step otherwise relatively weak reducing agent – glutathione.
of the process is the dissociation DMBI, followed by rapid com- Synthetic derivatives of cobalamin are known (so-called “Antivi-
plexation by the reducing agent, and electron transfer (Scheme 3). tamins B12 ”), which can minimize deligation at the Co within
Cyanide binds weakly to the Co(II) corrins and, therefore, leaves the the CblC-protein. Such complexes include 4-ethylphenylcobalamin
complex [70]. The reaction product in this case is Cbl(II)–SO2 •− , as (with a Co C(phenyl) bond) [89], and phenylethynylcobalamin
in the case of the aqua complexes. (with the acetylenic moiety bound at Co) [90]. Impaired deligation
Despite the very low value of the reduction potential of in vivo implies increased resistance toward transformation toward
Cbl(III)–CN− and the extremely low rate of its interaction with the coenzyme form, so that the enzymes binding these “antivita-
reducing agents, in living organisms, there are effective ways to mins” are deprived of their catalytic function. This is important, for
transform this complex to coenzyme forms. One of the stages of example, for the development of drugs that suppress cancer cells,
this process is reduction to Cbl(II). In mammals, decyanation of as these require large amounts of cobalamin for their growth. In
Cbl(III)–CN− is accomplished the CblC trafficking protein, which addition, inhibition of Cbl-dependent enzymes in vivo may be rele-
uses as the electron source methionine synthase reductase [75], or vant for a set of degenerative brain and nerve diseases, attributable
glutathione [76]. Binding of cyanocobalamin to the CblC protein to a lack of vitamin B12 .
leads to a base-off form [77,78], which is expected to facilitates the Thus, two types of mechanisms have been discussed for reduc-
decyanation [77,78] in ways reminiscent of the above-discussed tion of alkylcobalamins (Scheme 4). One type of mechanism
reduction of Cbl(III)–CN− by sulfur-based reductants [73]. involves the outer-sphere reaction of the reducing agent with a
There are several routes for chemical reduction of alkylkobal- base-off form of the alkylcobalamin, yielding (according to quan-
amins [79]. For example, carbon monoxide, sodium dithionite tum chemical calculations) a corrin anion radical [91]. The reduced
and sodium stannite reduce methylcobalamin in strongly alkaline form of the complex displays a dramatically increased rate of cleav-
media. The reaction proceeds at a low rate, but the removal of DMBI age of the Co C bonds (∼1015 times), leading to Co(I) and the alkyl
leads to a further increase in the reaction rate, of about 10-fold; radical of [92] or, in the presence of acids, Co(II) and the respective
indeed, the base-off forms of organocobalamins have a higher redox hydrocarbon [93]. Indeed, alkyl radicals were demonstrated using
potential than base-on ones [80]. spin trapping during electrochemical reduction of organocobal-
The interaction of alkylcorrinoids with thiols [81,82] pro- amins in DMF [94,95]. Decomposition of the reduced alkyl complex
ceeds at low rates, yielding cobalamin(II) and the corresponding may be facilitated by the transition to a base-off state [91]. In the
thioether. The mechanism involves attack of thiolate on the Co C second type of reduction mechanism, the reductants are thiols,

Scheme 3. A mechanism of reduction of cyanocobalamin.


Source: Adapted from Ref. [73].
74 I.A. Dereven’kov et al. / Coordination Chemistry Reviews 309 (2016) 68–83

Scheme 4. Mechanisms of reduction of alkylcobalamins


Source: Adapted from Refs. [91–97].

and they directly attack the Co C bond. The transition state in this Table 1
Thermodynamic and kinetic parameters of Cbl(II) ↔ Cbl(I) transformations at 22 ◦ C
case (RS-Me-Cbl) can be described either as closed-shell [96] or as
(adapted from [100]).
open-shell biradical in which one of the electrons has moved from
thiolate onto the corrin [91,97]. Reaction Value

pKa
4. Mechanisms of reduction reactions for Co(II) forms
(5) 2.9
Reduction of Cbl(II) to Cbl(I) proceeds much more compli-
cated than reduction of Cbl(III) to Cbl(II). The potential for the
Cbl(II)base-on /Cbl(I) transition is −0.61 V (vs. NHE, 22 ◦ C), while
under similar conditions the potential for the Cbl(II)base-off /Cbl(I) (6) 4.9a
pair is −0.50 V [32]. These values are significantly lower than the
potential of the biological reducing agent – generally considered
to be the flavin enzyme, methionine synthase reductase [98], for
which the potential of the semiquinone/hydroquinone is −0.28 V
(7) 1.0
(pH 7.0, 25 ◦ C) [99]. Values of pKa , standard potentials and rate
constants related to Cbl(II)/Cbl(I) species (reactions (5)–(11)) are
summarized in Table 1. Standard potentials (NHE), V
The set of reductants able to afford Co(I) corrinoids is rela-
tively small, including Zn [27], Ti(III) [101], NaBH4 [36], SO2 2−
−0.496
and HOCH2 SO2 − (both in alkaline media) [66,67], monosaccharides (8)

[35] and thiols [102,103] (the latter two only for Cbi(II) in alkaline
media).
The kinetics of the reaction between Cbi(II) and monosaccha-
rides [35] (notably, Cbl(II) is inert in such reactions) displays (9) −0.498
nonlinear concentration dependences which can be linearized in
inverse coordinates, and the reaction order for the hydroxide ion is
two. The mechanism of the reaction includes rapid reversible coor-
dination of the ionized form of the monosaccharide at the Co(II) (10) −0.607
center, leading to the formation of a pentacoordinated complex
(reactions (12) and (13)), and then the interaction of this complex Rate constants, s−1
with OH− with the subsequent transfer of the electron to Co(II)
(reaction (14)). Probably, the coordination of the reducing agent is (11) 105 (on-reaction)
not possible on Cbl(II), due to the trans DMBI. Reduction of Cbi(II) 1.6 × 103 (off-reaction)
proceeds well with thiols under basic conditions or in the presence a
pKa = 5.56 at 25 ◦ C [36].
of base in methanol, which is not observed for Cbl(II) [102].
GlcH ↔ Glc− + H+ . (12)
− − Interaction of Co(II)-corrinoids with stronger reducing agents
H2 O–Cbi(II) + Glc ↔ Glc–Cbi(II) + H2 O. (13)
(sulfoxylate, hydroxymethanesulfinate) proceeds well under strin-
gent conditions (strongly alkaline medium, an excess of reducing
− agent). The reduction of Co(II)–SO2 •− complexes of cobalamin and
Glc–Cbi(II) + OH−
cobinamide by sulfoxylate occurs with approximately equal rates:
→ Cbi(I) + products of oxidation of monosaccharide. (14) the rate constant for the reaction with Cbl(II)–SO2 •− is (2.83 ± 0.11)
I.A. Dereven’kov et al. / Coordination Chemistry Reviews 309 (2016) 68–83 75

Scheme 5. A mechanism of reduction of Co(II)-corrins by sulfoxylate (SO2 2− ).


Source: Adapted from Ref. [67].

M−1 s−1 vs. (2.18 ± 0.07) M−1 s−1 for Cbi(II)–SO2 •− at pH 13.0, 25 ◦ C Photosensitizers can facilitate the formation of Co(I)-corrinoids
[67]. The pH dependence of the process is described by an S- (Scheme 6). It is reported that in the presence of [Ru(II)(bpy)3 ]Cl2 ,
shaped curve with a pKa = 13.47 ± 0.07 at 25 ◦ C [67], consistent with the Co(I) form of heptamethylcobyrinate can be obtained by
deprotonation of the anion SO2 H− [104] and implying that the irradiation with visible light (wavelength not specified) [106]. Pho-
reducing agent for Co(II) is SO2 2− . In cases where a separate pre- tosynthesis of styrene-based polymers with covalently attached
liminary reduction to Co(II) is achieved using ascorbate (to avoid dicyanohexamethylcobyrinate and Ru(bpy)3 , using light at 
accumulation of the SO2 •− radical anion) allows for more facile ≥420 nm, has been reported to also enatil Co(I) formation
subsequent reduction to Co(I). This confirms that a Co(II)-anion- [107]. Simultaneous immobilization of heptamethylcobyrinate and
radical complex is not a prerequisite for reduction of Co(II) by SO2 2− Ru(bpy)3 in MOF (metal-organic frameworks) also allows Co(I)
(Scheme 5). formation when irradiated with visible light ( ≥ 420 nm) [108].
The addition of dithionite impairs conversion of Co(II) to Co(I), Alternative photosensitizers in this reaction may also be the less
and at [S2 O4 2− ] > 0.05 M Co(I) is no longer obtained, which indi- expensive dyes Rose Bengal [109] and Rhodamine B [110] – using
cates the occurrence of the reverse reaction, oxidation of Co(I) by  ≥ 440 nm.
dithionite. Indeed, addition of dithionite to Co(I) corrins leads to Co(II)-corrinoids may also be photoreduced using TiO2
their oxidation to Co(II)–SO2 •− [66]. [111,112]. Cyanocobyrinic acid applied to a TiO2 surface is reduced
For the reaction of Co(II) complexes with hydroxymethanesul- to Co(I) upon UV irradiation ( = 365 nm) [112].
finate (HMS), the dependence of the observed rate constant on As shown above, the reduction of cobalamin(II) complexes is a
the concentration of HMS is linear only in the inverse coordi- relatively challenging reaction; nevertheless, living organisms have
nates (1/kapp. vs. 1/[HMS]), which is typical of a reaction occurring developed catalytic mechanisms to implement this process. One
through rapid reversible complex formation and subsequent decay. such pathway entails conversion of the Cbl(II) to the tetracoordi-
The observed reaction rate constant increases linearly with [OH− ], nated form upon binding to the enzyme active site – as it is indeed
indicating that hydroxide participates in the process. Thus, the for- expected that the removal of nucleophiles from the coordination
mation of Co(I) entails reactions (15) and (16) [67]. sphere will reduce the potential of the Co(II)/Co(I) couple [27]. Such
a complex was detected as part of the ATP: cobalamin adenosyl-
Co(II)–SO2 •− + HOCH2 SO2 − ↔ − HOCH2 SO2 –Co(II) + SO2 •− . (15) transferase [24–26,113]. An alternative route for the reduction of
cobalamin(II) is based on the hypothesis of the formation of Co(I)
– H–X hydrogen bonds [28–31]. The DFT-calculated value of such
− bond lengths would be ∼2.25 Å, depending primarily on the type
HOCH2 SO2 –Co(II) + OH−
of substrate [28]. Calculations show that the transition from the
→ Co(I) + products of oxidation of HMS. (16) pentacoordinated Cbl(II)base-off to pseudo-pentacoordinated Cbl(I)
– H–X is thermodynamically more favorable than the transition to a
simple tetracoordinated Cbl(I), and the process is even further facil-
Kinetics data suggest a dependence of the reduction mecha- itated if two hydrogen bonds are involved instead of one [28,29].
nisms of Co(II) cobalamins and cobinamides on the nature of the To date, it has not been possible to obtain crystallographic informa-
reducing agent. Thus, reduction of Cbl(II) and Cbi(II) with sulfoxy- tion on Cbl(I), presumably due to its unusually high reactivity. In
late or sodium hydroxymethanesulfinate (strong electron donors) favor of this hypothesis is the fact that hydrogen bonding to simpler
proceeds with comparable rates, while reduction of the same metal (non-corrinoid) Co(I) centers is known experimentally [114].
complexes by monosaccharides (weaker electron donors) proceeds Activation of the corrinoid-[Fe–S] protein, which carries out the
differently. transfer of a methyl group from methyltetrahydrofolate to acetyl-
Co(I) corrinoids may be prepared photochemically by irra- CoA, is also instructive in these respects. Binding of the activator (an
diation in solution at appropriate wavelengths. For example, iron-sulfur protein itself) is essential for electron transfer from the
without added photosensitizer the Co(I) heptametilcobyrinate [2Fe2S] cluster to the Co(II)-corrinoid in the presence of ATP. This
can be obtained by irradiation of a methanolic solution or binding leads to significant structural changes, which are supported
a solution in an ionic liquid (N-methyl-N-propylpyrrolidinium by resonance Raman [115] and EPR spectroscopy [115,116] as well
bis-(trifluoromethanesulfonyl) amide) in the presence of tri- as by crystallographic data [116]. Thus, a water molecule (∼2.5 Å
ethanolamine, with 365-nm light [105]. In the ionic liquid the Co(II) – O bond length [117]) at the Cbl ␤-position is replaced
reaction is ∼10 times faster. Two mechanisms have been proposed (upon binding of the activator protein) by a serine HO side-
for this process. In the first case, corrin photoexcitation leads to chain with a Co(II) – H contact at 2.5 Å [115]. The reason for this
electron transfer from the macrocycle to metal, thereby forming a replacement is unclear (the authors of [116] suggest that this sub-
corrin radical cation. In the second case, a transient complex may be stitution contributes to the displacement of the redox potential of
formed between the photoexcited corrinoids and triethanolamine, the Co(II)/Co(I) couple in the negative direction, i.e., stabilizes the
allowing electron transfer to the cobalt (Scheme 6). Co(II) state).
76 I.A. Dereven’kov et al. / Coordination Chemistry Reviews 309 (2016) 68–83

Scheme 6. Possible mechanisms of UV–Vis light driven Cbx(II) reduction without (A) or with (B) photosensitizer added in system.
Source: Adapted from Refs. [105,110].

5. Mechanisms of oxidation of the Co(II) forms of Probably the most thoroughly investigated reaction of Cbl(II)
cobalamins and their derivatives is the one with NO. This reaction can take place in vivo, as NO
was shown to be able to inhibit Cbl-dependent enzymatic pro-
Despite its relatively high redox potential, Cbl(II) is a fairly cesses [128,129] and Cbls influence NO-related events [130–132].
strong reducing agent of free radicals. This is explained primarily by NO binding to cobalamin proceeds at a high rate (the rate constants
the presence of an unpaired electron in the metal 3dz2 orbital, per- at pH 1.0 and 7.4, 25 ◦ C, are 5.0 and 7.4 × 108 M−1 s−1 , respectively)
pendicular to the plane of the macrocycle. Contact with an unpaired and is reversible (the dissociation rate constants at pH 3.6 and 7.4,
electron of the oxidant hardly affects macrocycle and minimizes 25 ◦ C are 1.7 and 5.6 s−1 , respectively) [133]. The resulting complex
destruction of the complex during the reaction [118]. is EPR-silent [133], its resonance Raman spectrum shows band at
Co(II)-corrinoids react with oxygen through a fast reversible 514 cm−1 assigned to a bent Co(III)–NO− bond [134]. Nitrogen(II)
oxygen binding step, eventually yielding the Co(III) state (reaction within a nitroxyl NO− ligand was confirmed later by 15 N NMR [133]
(17)). Evidence for a transient Co(III)–O2 •− adduct has come form and X-ray analysis [135,136]. DFT calculations show, however, that
EPR [119–122], quantum chemical calculations [123], and crystal- a more precise description of the structure of the complex would
lography [124]. be a hybrid between the Cbl(III)–NO− and Cbl(II)–NO resonance
structures [137].
Co(II) + O2 ↔ Co(III)–O2 •− →→ Co(III)–solv. + H2 O2 + O2 . (17) The activation parameters of the direct reaction at
solv. = H2 O, MeOH, etc. pH 7.4 are: H =/ = 24.5 kJ mol−1 , S =/ = +7 J K−1 mol−1 ,
Cbl(II) reacts very rapidly with superoxide [125,126], yield- V =/ = +5.4 cm3 mol−1 . Since the small positive values of entropy
ing Cbl(III) without modification of the corrin (reaction (18)). and volume activation are characteristic of a dissociative inter-
The rate constant (measured with a steady-state flux of change (Id ) mechanism, the authors proposed a scheme for the
superoxide from the xanthine oxidase/acetaldehyde system) is reaction with a rapid reversible formation of a hexacoordinated
(6.8 ± 0.8) × 108 M−1 s−1 at pH 7.4, 25 ◦ C, which is comparable to aqua-cobalamin(II) intermediate, and the subsequent replacement
the rate constant for superoxide dismutation (2 × 109 M−1 s−1 in of bound water molecule by NO (Scheme 7) [133].
the presence of Cu, Zn – superoxide dismutase) [125]. Similar rates The rate of the forward reaction between Cbi(II) and NO is
((3.78 ± 0.07) × 108 M−1 s−1 ) were observed when superoxide was comparable to that of Cbl(II) (2.4 ± 108 M−1 s−1 at pH 7.4, 25 ◦ C),
obtained by radiolysis. Co(III) corrinoids do not react with super- but the rate of reverse process is significantly lower (0.019 s−1 at
oxide [126]. pH 7.4, 25 ◦ C) [138]. The higher affinity of Cbi(II) to NO, and the
absence of a base in the ␣-position, lead to a significant differ-
Cbl(II) + O2 •− + 2H3 O+ → Cbl(III)–OH2 + H2 O2 + H2 O. (18) ence in the properties of the Cbl(II) and Cbi(II) nitrosyl complexes:
nitrosylcobinamide is relatively stable in the presence of oxygen
The reaction between Cbl(II) and H2 O2 occurs at a slower rate
[139], whereas nitrosylcobalamin rapidly oxidized to a mixture of
[125]. The H2 O2 is able to destroy the corrin [33]; this happens
nitro and aquacobalamin [140]. In this context, Cbi(II)–NO has been
in systems containing both reductants (e.g., ascorbic acid [33,39],
thiols [127], polyphenolic compound [127]) and molecular oxy-
gen. Under such conditions, Co(III) ↔ Co(II) redox cycling leads to
considerable amounts of H2 O2 as side-products. Then, according
to reaction (19), hydroxyl radicals may cause degradation of the
cobalamin. This phenomenon occurs in foods fortified with vita-
min B12 and containing a significant amount of reducing agents
such as polyphenols [127].
Scheme 7. A mechanism of nitric oxide binding by Cbl(II).
Cbl(II) + H2 O2 → Cbl(III)–OH− + OH• . (19) Source: Adapted from Ref. [133].
I.A. Dereven’kov et al. / Coordination Chemistry Reviews 309 (2016) 68–83 77

proposed as hypotensive agent [139] as well as wound healing instead reacts with NO (reaction (21)), so that the final products
accelerator [141]. In addition, cobinamide can act as an effective are Cbl(III)–NO− and Cbl(III)–OH2 .
NO trap in the body, with potential use for controlling the concen-
tration of this important signaling molecule [142]. HNO2 ↔ NO2 − + H+ . (27)
There are several routes toward Cbl(III)–NO− . For instance, 2HNO2 ↔ NO+ + NO2 − + H2 O. (28)
Cbl(III)–SG− reacts very rapidly with NO (2.8 × 103 M−1 s−1 at pH
7.0), yielding Cbl(III)–NO− via a mechanism proposed to involve NO+ + Cbl(II) → Cbl(III) + NO. (29)
formation of Cbl(II)–ONSG intermediate further decomposing to
Cbl(II) reacts rapidly with sodium nitroprusside
Cbl(III)–NO− and thyil radical [143]. The subsequently-derived
(Na2 [Fe(CN)5 NO]). In the initial step, there is a rapid reversible
activation parameters (H =/ = 41 kJ mol−1 , S =/ = –42 J K−1 mol−1 ,
complex formation between [Fe(II)(CN)5 NO]2− and Cbl(II) (reac-
V =/ = –9.6 cm3 mol−1 ) are characteristic for an associative mech-
tion (30)), which after intramolecular electron transfer becomes
anism [144] and support a transient formation of Cbl(II)–ONSG.
Cbl(III)–[Fe(I)(CN)4 NO]2− and CN− (reaction (31)). The latter is
Alternatively proposed mechanism for the reaction [144] involves
then able to substitute the [Fe(I)(CN)4 NO]2− fragment on Co(III)
two NO-molecules and proceeds via reversible formation of an
(reaction (32)). Nitroprusside is also capable of substituting the
intermediate complex of Cbl(III)–SG− with NO (as a rate determin-
[Fe(I)(CN)4 NO]2− fragment (reaction (33)) [149].
ing step), followed by fast decay to Cbl(II) and nitrosothiol (reaction
(20)). The excess NO shifts the equilibrium toward the products Cbl(II) + [Fe(II)(CN)5 NO]2− ↔ Cbl(II)–[Fe(II)(CN)5 NO]2− . (30)
(reaction (21)) [144]. Nevertheless, reaction stoichiometry was not
determined providing no preference for any proposed pathways.

Cbl(III)–SG− + NO ↔ Cbl(II)–ONSG ↔ Cbl(II) + GSNO. (20) Cbl(II)–[Fe(II)(CN)5 NO]2− ↔ Cbl(III)–[Fe(I)(CN)4 NO]2− + CN− .

(31)
Cbl(II) + NO ↔ Cbl(III)–NO− . (21)

Nitroxylcobalamin is formed during the reaction of aquacobal-


amin with R2 N–NONO derivatives (1-(N,N-dialkylamino)diazen-1-
ium-1,2-diolates). The mechanism of the reaction involves rapid Cbl(III)–[Fe(I)(CN)4 NO]2− + CN−
reversible coordination of R2 N–NONO to Co(III) and the subsequent
↔ Cbl(III)–CN− + [Fe(I)(CN)4 NO]2− . (32)
slow decay of the complex to Cbl(III)–NO− and the corresponding
nitrosamine (reaction (22)) [145].

Cbl(III)–OH2 + R 2 N–NONO ↔ R 2 N–NONO–Cbl(III) + H2 O


Cbl(III)–[Fe(I)(CN)4 NO]2− + [Fe(II)(CN)5 NO]2−
→ Cbl(III)–NO− + R 2 N–NO. (22)
↔ Cbl(III)–[Fe(II)(CN)5 NO]2− + [Fe(I)(CN)4 NO]2− . (33)

The authors of [146] believe that Cbl(III)–NO− is formed dur-


ing the reaction of Cbl(III)–OH2 with Angelli’s salt (Na2 N2 O3 ). The Cbl(II) reacts very rapidly with nitrogen dioxide (NO2 • ) (rate
reaction mechanism includes two parallel routes: direct interac- constant at pH 7.4, 25 ◦ C: 3.5 ± 108 M−1 s−1 ). The product (reaction
tion of Cbl(III)–OH2 with the decay product of Angelli’s salt (HNO, (34)) is nitrocobalamin, which does not further interact with NO2 •
reactions (23) and (24), noting that the side product NO2 − may [150].
slow down reaction (24)), and a rapid reversible complex for-
Cbl(II) + NO2 • → Cbl(III)–NO2 − . (34)
mation between Cbl(III) and HN2 O3 − with subsequent decay to
Cbl(III)–NO− and NO2 − (reaction (25)). Cbl(II) efficiently reacts with peroxynitrous acid (ONOOH): the
− − rate constant at 25 ◦ C is 1.6 × 106 M−1 s−1 . During the one-electron
HN2 O3 → HNO + NO2 . (23)
reduction ONOOH forms NO2 • (reaction (35)), which is then further

HNO + Cbl(III)–OH2 → Cbl(III)–NO + H2 O + H . +
(24) reduced to NO2 − (reaction (34)) [151].

Cbl(II) + ONOOH → Cbl(III)–OH− + NO2 • . (35)

Cbl(II) reacts with nitroxyl (HNO/NO− )


yielding Cbl(III)–NO−
Cbl(III)–OH2 + HN2 O3 − ↔ Cbl(III)–HN2 O3 − + H2 O
and N2 as final products. Reaction stoichiometry was found to
→ Cbl(III)–NO− + NO2 − . (25) be [Cbl(II)]0 : [HNO]0 = 1: 2 and authors suggested the following
mechanism (reactions (21), (36) and (37)) [152]. In this scheme,
generation of Cbl(I) by reaction (36) seems to be disadvantageous,
Generation of Cbl(III)–NO− and Cbi(II)–NO may also occur by since Cbl(I) is very reactive toward NO formed (see below). DFT
reductive nitrosylation of Cbl(III) with a mixture of NO and NO2 − and CASSCF calculations predict an existence a Cbl(II)–HNO adduct
at pH <4. The process is reversible and can be represented as [123], which can be the first intermediate in this mechanism.
the macroscopic equilibrium (reaction (26)) (K = 0.6 ± 0.2 at 25 ◦ C)
[147]: Cbl(II) + HNO → Cbl(I) + NO + H+ . (36)

Cbl(III)–NO2 − + 2NO + H2 O ↔ Cbl(III)–NO− + 2HNO2 . (26) 2Cbl(I) + 2HNO + 2H+ → Cbl(I) + N2 + H2 O. (37)

Also very fast is the reaction between Cbl(II) and the carbonate
Cbl(II) reacts with nitrite under acidic conditions (reactions
radical (CO3 •− ). It consists of three successive stages. During the
(27)–(29)). The rate-determining step is nitrosonium ion (NO+ )
first stage (rate constant at pH 9.0: 2.0 × 109 M−1 s−1 ; reaction (38))
formation from with nitrous acid (28). The subsequent reaction
Cbl(II) is oxidized to hydroxocobalamin, while the subsequent two
between Cbl(II) and NO+ (reaction (29)) occurs at a higher rate
stages entail modifications of the corrin ring [153].
and leads to the formation of Cbl(III) and NO, which do not further
react with each other [148]. The remaining Cbl(II) in the system Cbl(II) + CO3 •− + H2 O → Cbl(III)–OH− + HCO3 − . (38)
78 I.A. Dereven’kov et al. / Coordination Chemistry Reviews 309 (2016) 68–83

Based on the above data it can be concluded that Co(II)- potential of the Co(III)/Co(II) couples, i.e. increases the reducing
corrinoids effectively react with many species responsible for properties of cobalamin(II) [155] as yet another illustration of the
oxidative as well as nitrosative stress. Not surprisingly then, in vitro manner whereby the coordination environment of the ion Co(II)
experiments have shown that cobalamins have antioxidant prop- directly modulates its redox properties. Perhaps further study of
erties [154]. redox reactions involving other Co(II)-corrinoids may allow more
One feature of the Co(II)-corrinoids is their ability to methylate insight, as their formation occurs much more easily than that of
with methyl iodide [102,103], and in some cases, methyl tosylate Cbl(II) complexes [158].
[102] in the presence of thiols. It is important to stress that there is Cbl(II) reacts with inorganic Co(III) complexes (chloride, bro-
no direct reaction between Cbl(II) and the methylating agent. Two mide, iodide, azide, thiocyanate, or pyrazine derivatives of
mechanisms were proposed for these processes, depending on the Co(III)(NH3 )5 and Co(III)(bpy)3 ). The reaction products are aqua-
pH, the type of the metal complex, thiol and methylating agent. cobalamin and the Co(II) salt. The electron transfer occurs in
The first is due to the ability of thiols to reduce of Co(II) base-off inner-sphere manner with Co(III)(NH3 )5 , while for Co(III)(bpy)3 it
corrinoids in alkaline media to Co(I) (reaction (39)), which readily is outer-sphere [159].
accepts a methyl group (reaction (40)). The second route is observed
in acidic environments, reduction of Co(II) is impossible at low con-
centrations of thiolate; here, the reaction is assumed to involve a 6. Oxidation mechanisms for Co(I) cobalamins and their
Co(II)–RSH complex (reaction (41)), which then reacts with MeI derivatives
forming a so-called ‘Pratt complex’ [RSH–Co(II)–MeI], which fur-
ther transforms into Me–Co(III) (reaction (42)) [102,103]. Cbl(I) is a very strong natural nucleophile (“supernucleophile”)
and a strong reductant. On the Pearson scale it displays a nucle-
2Co(II) + 2RSH ↔ 2Co(I) + RSSR + 2H+ . (39)
ophilicity of 14.4 [160,161], which is much higher than of other
Co(I) + MeI →→ Me–Co(III) + I− . (40) biological centers. According to CASPT2 [162] and CASSCF [163]
calculations, the Cbl(I) ground state entails a mixture of Co(I) (d8 )
Co(II) + RSH ↔ RSH–Co(II). (41) and Co(II)-(d7 )-corrin radical configurations, with a significant pre-
dominance of the former. The reason for the high reactivity of Cbl(I)
can be a significant destabilization of the 3dz2 orbital and its conve-
RSH–Co(II) + MeI ↔ RSH–Co(II)–MeI nient orientation for contact with an oxidizing agent [27], as well as
the presence of the biradical component in the ground-state wave
→ Me–Co(III) + ½RSSR + I− . (42)
function [163].
One of the most prominent reactions of Co(I) complexes is the
While Cbl(II) does not directly react with dehydroascorbic acid one with protons – and Co(I)-corrinoids are not exception: they
(DHA), the addition of GSH [155], cysteine [156], N-acetylcysteine are poorly stable in aqueous solutions, eventually yielding Cbl(II)
[156] or SCN− [155] leads to oxidative formation of RS− –Cbl(III) and H2 in an process displaying oscillatory kinetics. Oscillations are
complexes, and two-electron reduction of DHA to ascorbic acid. observed in the pH range 2–12; their period depends on the tem-
Importantly, this reaction does not proceed in the presence of other perature, and they disappear under the action of light. However,
amino acids (those that do not contain sulfur, and even methionine their origin is unclear and no reaction intermediates were observed
[156]), while in the presence of S-methylglutathione the process experimentally [164].
proceeds with a very low rate [155]. Detailed kinetic data are avail- Cbl(I) is much more stable in acidic environments than other
able for this reaction in acidic and neutral media, where the rate Co(I) complexes, and displays a protonation pKa of ∼1 (reaction
increases with pH, and with [RSH]. The mechanism involves rapid (7)) [165], much smaller than for other Co(I) complexes [166].
reversible coordination of the thiol to Co(II); interaction of this com- Hydrogen-generating photocatalytic systems based on corri-
plex with DHA leads to RS− –Co(III) and ascorbyl radical – which noids have been proposed [112,166]. The yield of hydrogen by
dismutates to dehydroascorbic and ascorbic acid (Scheme 8) [156]. irradiating a neutral Cbi aqueous solution in the presence of tri-
The formation of Cbl(II) complex with SCN− was confirmed exper- ethanolamine or eosin Y, was lower than using cobaloxime as
imentally by EPR and UV–Vis spectroscopy, as well as by quantum catalyst [166]. Significantly higher yields of hydrogen were afforded
chemical calculations [23], while the GSH–Cbl(II) complex was by cyanoaquacobyrinate immobilized on TiO2 . In the latter case,
confirmed by cryo-MS [155] and EPR [157], which is, probably, pen- addition of spin-trapping compounds affects the hydrogen yield,
tacoordinated [157]. Further quantum chemical calculations can which suggests involvement of H-radicals in the process [112].
provide more precise insight in structure of Co(II)-complexes with The proposed mechanism for the interaction of Cbl(I) with
thiols relevant to this reaction. Such complexation lowers the redox protons [112] involves, as a key step, the formation of hydridocobal-

Scheme 8. A mechanism of thiolate-assisted reduction of DHA by Cbl(II).


Source: Adapted from Ref. [156].
I.A. Dereven’kov et al. / Coordination Chemistry Reviews 309 (2016) 68–83 79

amin(III) (reaction (43)), which then dismutates, or reacts with Cbl(III)–H− + N2 O → Cbl(III)–OH− + N2 . (57)
protons (directly or after prior reduction to Co(II)-hydride), yielding
Reduction of hydroxylamine occurs at relatively low rate [175].
molecular hydrogen (reactions (44)–(47)). Hydridocobalamin was
The active form of the oxidant in this case is NH3 OH+ , probably
also assumed as a reagent reduction of double bonds [112].
according to reaction (58). Methylation of the nitrogen atom of
Co(I) + H+ → Co(III)–H− . (43) hydroxylamine increases the rate of reaction.

2Co(III)–H → 2Co(II) + H2 . (44) NH3 OH+ + Cbl(I) → NH3 + Cbl(III)–OH− . (58)
− + •−
Co(III)–H + H → Co(III) + H2 . (45) Sulfite reacts with Cbl(I) to yield a Cbl(II)–SO2 complex, in
− −
a process where HSO3 − , S2 O5 2− and SO2 ·H2 O may act as oxi-
Co(III)–H + ē → Co(II)–H . (46) dants (reactions (59)–(61)), wherein the reactivity of the substrates
− +
Co(II)–H + H → Co(II) + H2 . (47) decreases in the series: SO2 ·H2 O > S2 O5 2− > HSO3 − . Very rapid is
also the reaction of Cbl(I) with thiosulfate. In this case, the oxidant
Cbl(I) also reacts with nitrogen oxides and oxyanions, such as is probably HS2 O3 − (reaction (62)), which is ultimately converted
NO3 − , ONOO− , NO2 − , NO, N2 O, NH2 OH. The reaction with nitrate to HS− and SO3 2− (reaction (63)) [176].
[167,168] in acidic medium has NH4 + as final product [168]; the
active form reacting with Cbl(I) is HNO3 (reaction (48)) [168]. Cbl(I) Cbl(I) + HSO3 − → Cbl(II)–SO2 •− + OH− . (59)
reacts with HNO2 , eventually leading to the formation of NH2 OH in Cbl(I) + SO2 ·H2 O → Cbl(II)–SO2 •− + H2 O. (60)
slightly alkaline media (reactions (49) and (50)) [168]. It is likely
Cbl(I) + S2 O5 2−
→ Cbl(II)–SO2 •− + SO3 2−
. (61)
that nitrite is an intermediate product of nitrate reduction, and
variations in the composition of the final reaction products may be
explained by the pH dependence of the reactivity of hydroxylamine
(see below). Cbl(I) + HS2 O3 − → Cbl(II) + HS• (or SO3 •− ) + SO3 2− (or HS− ).

Cbl(I) + HNO3 → Cbl(III)–OH− + NO2 − . (48) (62)


+ −
Cbl(I) + HNO2 + H → Cbl(III)–OH + HNO. (49)
+ −
Cbl(I) + HNO + H + H2 O → Cbl(III)–OH + NH2 OH. (50) Cbl(I) + HS• (or SO3 •− ) → Cbl(II) + HS− (or SO3 2− ). (63)

Cbl(I) reacts very rapidly with peroxynitrite [169] (in its ONOO− The reaction of Cbl(I) with dithionite is the only known example
as well as ONOOH forms), with N2 reported as product. It is not of chemical reduction of this compound. In this process Cbl(I) reacts
clear what is the difference in the mechanisms of the reduction with dithionite itself, S2 O4 2− , as well as with its monomerization
of the two isomers (nitrate and peroxynitrite), leading to different product, SO2 •− . Since sulfoxylate is formed during the reaction, and
final products, but one can assume that a key role is played by the it can act as reducing agent for Cbl(II), the following macroscopic
activation of ON OOH bond, which leads to the formation of NO – equilibria can be proposed (reactions (64) and (65)) [176].
an intermediate not common in the reduction of HNO3 . Cbl(I) + SO2 •− ↔ Cbl(II) + SO2 2− . (64)
Cbl(I) reacts rapidly with NO in solution, yielding to N2 O and
Cbl(I) + S2 O4 2−
↔ Cbl(II)–SO2 •− + SO2 2−
. (65)
N2 [170]. On the other hand, electrochemical reduction of NO on a
glassy carbon electrode modified with cyanocobalamin, leads to Cbl(I) is capable of reducing disulfides to thiols. Possibly, dur-
NH3 and NH2 OH [171]. Under such conditions, cyanocobalamin ing the rate-determining step (reaction (66)), a mixture of thiol
is reduced to Cbl(I). Probably, HNO is produced in the course of and thiyl radical is formed–with the latter affording a more rapid
the one-electron reduction of NO (reaction (52)), followed by rapid subsequent reaction with Cbl(I) (reaction (67)) [177].
dimerization and decomposition to N2 O (reaction (52)).
Cbl(I) + RSSR → Cbl(II) + RS− + RS• . (66)
Cbl(I) + NO + H+ → Cbl(II) + HNO. (51)
Cbl(I) + RS• → Cbl(II) + RS− . (67)
2HNO ↔ H2 N2 O2 → N2 O + H2 O. (52)
Cbl(I) also reduces various transition metal ions. Thus, it reduces
Cbl(I) is capable to reduce HNO [152]. In this case, neither V(IV) (vanadyl) and Mo(VI) (molybdate) to V(II) and Mo(III), respec-
NH4 + nor NH2 OH are reaction products, which implies reduction of tively (reactions (68)–(71)), and this reaction is autocatalytic, which
HNO to N2 . Authors suggested the following mechanism (reactions is explained by the higher oxidizing properties of V(III) and Mo(V),
(53) and (54)) involving transient formation of iminoxyl radical formed in the initial one-electron reduction [178].
(H2 NO• ). Therefore, reactions (51)–(54) can explain the formation
of N2 O and N2 as final products of reaction between Cbl(I) and NO. Cbl(I) + V(IV) → Cbl(II) + V(III). (68)
+
Cbl(I) + HNO + H → Cbl(II) + H2 NO• . (53) Cbl(I) + V(III) → Cbl(II) + V(II). (69)

2H2 NO• → N2 + 2H2 O. (54) Cbl(I) + Mo(VI) → Cbl(II) + Mo(V). (70)

It is also possible to explain a number of products of reactions Cbl(I) + Mo(V) → Cbl(II) + Mo(IV). (71)
between Cbl(I) and nitrogen oxy-derivatives by the reason in the It is known that Cbl(I) exhibits high reactivity toward various
difference in the coordination mode of NO-motif (nitro vs. nitrito), organic compounds: methyl group donors [179–190], epox-
as in the case of other tetrapyrrole complexes [172,173]. ides [191–193], halides [194–202], and unsaturated compounds
N2 O is reduced by Cbl(I) to N2 [41,174]. Two routes are proposed [193,203].
for this reaction, depending on the species acting as actual reduc- Interaction of Cbl(I) with CH3 -donors is a key step in the
tant: Cbl(I) directly (reactions (55) and (56)) or hydridocobalamin catalytic cycle of the transfer of a methyl group in living organ-
(reaction (57)) [41]. isms. Such donors may be methyltetrahydrofolate [101,179],
Cbl(I) + N2 O → Cbl(II) + N2 O− . (55) methylamines [180–182], methyl sulfides [183], methoxides [184],
methyl phosphates [185]. Notably, the direct interaction of Cbl(I)
− + −
Cbl(I) + N2 O + H → Cbl(II) + N2 + OH . (56) with CH3 -donors takes place at very low rates [186], and it remains
80 I.A. Dereven’kov et al. / Coordination Chemistry Reviews 309 (2016) 68–83

Scheme 9. A mechanism of Cbl(I)-assisted dehalogenation of DDT.


Source: Adapted from Refs. [94,200].

unclear how enzymes achieve acceleration of this reaction. Tra- the main products are 1,1-bis(4-chlorophenyl)-2,2-dichloroethane
ditionally, the process is assumed to entail an SN2 -mechanism (DDD), 1,1-bis(4-chlorophenyl)-2,2-dichloroethylene (DDE),
of nucleophilic substitution, but one cannot exclude an alterna- 1-chloro-2,2-bis(4-chlorophenyl)ethylene (DDMU) and 1,1,4,4-
tive mechanism involving electron transfer, where the reaction tetrakis(4-chlorophenyl)-2,3-dichloro-2-butene (TTDB) [94],
proceeds via a Co(II) – pterin–radical complex, and a subsequent while in ionic liquids 1,19-(ethylidene)-bis(4-chlorobenzene)
transfer of the methyl group [187–190]. (DDO), 1,19-(ethenylidene)-bis(4-chlorobenzene) (DDNU) and
Epoxides (oxiranes) are toxic metabolites produced, among oth- 1,19-(2-chloroethylidene)-bis(4-chlorobenzene) (DDMS) are
ers, by cytochrome P450 enzymes acting on compounds containing reported [200]. The proposed mechanism involves nucleophilic
double bonds (acrylamide, 1,3-butadiene and others). With Cbl(I), substitution of the first chlorine atom in DDT by cobalamin(I),
epoxides form a stable complex via a Co C bond. Such processes yielding an organocobalamin. Subsequent cleavage of the Co C
have been employed in proposed efficient methods for the deter- bond by light or electrolysis yields an organic radical, which when
mination of epoxides in vivo [191–193]. liberated into the solvent is converted into the above range of
Cbl(I) effects dehalogenation reactions of organic compounds, products (Scheme 9) [94]. Also studied has been the reaction
and has as such been proposed to be involved in forms of anaerobic between Cbl(I) and sucralose (Suc), an artificial sweetener con-
respiration, where the final electron acceptor is a halogenated taining three chlorine atoms. An initial substitution of one of the
compound (“halorespiration”). Such dehalogenations are also three chlorine atoms leads to the corresponding organocobalamin
of ecological interest, considering the halogenated nature of (Cbl–Suc), which may also possibly occur in vivo [201]. This com-
many organic environmental pollutants (e.g., DDT, polychlorethy- plex can collapse under the influence of light, acids, or electrolysis
lene). Although for polychlorethylene the reaction is reported [202].
in several instances [194–198], there is no clear understand- The dehalogenation processes of chlorlorethylene in vivo and
ing of the reaction mechanism. Most recently [198,204] it has in vitro are generally similar to each other [205], with differences
been proposed that the interaction of tetrachlorethylene or discussed below. During the first stage of the catalytic cycle of
trichlorethylene with Cbl(I) leads to the formation of solvent- the reductive dehalogenase enzyme, outer-sphere electron transfer
caged [Cbl (II)–vinyl• ] radical pairs, which can then either lead occurs from Co(I) to the substrate (over a distance of ∼5.8 Å [206]),
to an organocobalamin, or the vinyl radical may be liberated which leads to the formation of a halogen ion and an organic rad-
into the solvent. A variety of methods have been developed ical [206,207]. A direct interaction of the Co(II) with the radical
for the dehalogenation of 1,1-bis(4-chlorophenyl)-2,2,2- is in principle possible, leading to an organocobalamin (and more
trichloroethane (DDT) [94,105,106,108–110,199,200]. In DMF, so if the reaction occurs with free Cbl(I) instead of the enzyme).
I.A. Dereven’kov et al. / Coordination Chemistry Reviews 309 (2016) 68–83 81

Acknowledgements

This work was supported by Russian Scientific Foundation,


agreement No. 14-23-00204, Council on Grants of the President
of the Russian Federation for state support of young Russian
researchers (project number MK-3661.2015.3), and the Romanian
Ministry of Education and Research (PCE 418/2012).

References

[1] K.L. Brown, Chem. Rev. 105 (2005) 2075–2149.


[2] R. Banerjee (Ed.), Chemistry and Biochemistry of B12 , Wiley, New York, 1999.
[3] B. Kräutler, Biochem. Soc. Trans. 33 (2005) 806–810.
[4] K.Ó. Proinsias, M. Giedyk, D. Gryko, Chem. Soc. Rev. 42 (2013) 6605–6619.
[5] F. Zelder, K. Zhou, M. Sonnay, Dalton Trans. 42 (2013) 854–862.
[6] R.G. Matthews, Met. Ions Life Sci. 6 (2009) 53–114.
[7] R. Banerjee, ACS Chem. Biol. 1 (2006) 149–159.
[8] R. Banerjee, C. Gherasim, D. Padovani, Curr. Opin. Chem. Biol. 13 (2009)
484–491.
[9] C. Gherasim, M. Lofgren, R. Banerjee, J. Biol. Chem. 288 (2013) 13186–13193.
[10] K.L. Brown, Dalton Trans. (2006) 1123–1133.
[11] M. Giedyk, K. Goliszewska, D. Gryko, Chem. Soc. Rev. 44 (2015) 3391–3404.
[12] B. Kräutler, Chem. Eur. J. 21 (2015) 11280–11287.
[13] F. Zelder, M. Sonnay, L. Prieto, ChemBioChem 16 (2015) 1264–1278.
[14] F. Zelder, Chem. Commun. 51 (2015) 14004–14007.
[15] H. Shimakoshi, Y. Hisaeda, TCIMAIL (2009) 2–11.
Scheme 10. A mechanism implicated in reductive dehalogenase cycle toward [16] Y. Hisaeda, T. Nishioka, Y. Inoue, K. Asada, T. Hayashi, Coord. Chem. Rev. 198
debromination of ortho-dibromophenol. (2000) 21–37.
[17] M.S.A. Hamza, R. van Eldik, P.L.S. Harper, J.M. Pratt, E.A. Betterton, Eur. J. Inorg.
Source: Adapted from Ref. [208].
Chem. (2002) 580–583.
[18] H.A. Hassanin, L. Hannibal, D.W. Jacobsen, K.L. Brown, H.M. Marques, N.E.
Brasch, Dalton Trans. (2009) 424–433.
[19] B. Kräutler, W. Keller, C. Kratky, J. Am. Chem. Soc. 111 (1989) 8936–8938.
The organic radical then receives a second electron either from an [20] S. Van Doorslaer, G. Jeschke, B. Epel, D. Goldfarb, R.-A. Eichel, B. Kräutler, A.
FeS-cluster, or from Co(II). Differences were reported between the Schweiger, J. Am. Chem. Soc. 125 (2003) 5915–5927.
enzymatic dehalogenation of halogenophenols, versus processes [21] M. Giorgetti, I. Ascone, M. Berrettoni, P. Conti, S. Zamponi, R. Marassi, J. Biol.
Inorg. Chem. 5 (2000) 156–166.
involving free Cbl(I); the enzymatic reaction leads to the forma- [22] D.S. Salnikov, R. Silaghi-Dumitrescu, S.V. Makarov, R. van Eldik, G.R. Boss,
tion of a halide-cobalamin(III) complex (Scheme 10) instead of Dalton Trans. 40 (2011) 9831–9834.
organocobalamin [208], suggesting no Co C bond is involved in [23] I.A. Dereven’kov, D.S. Salnikov, S.V. Makarov, M. Surducan, R. Silaghi-
Dumitrescu, G.R. Boss, J. Inorg. Biochem. 125 (2013) 32–39.
the course of reductive dehalogenation in vivo.
[24] T.A. Stich, M. Yamanishi, R. Banerjee, T.C. Brunold, J. Am. Chem. Soc. 127 (2005)
The Cbl(I) reactions with alkenes [203] and alkynes [196] lead to 7660–7661.
organocobalamins containing alkane and alkene fragments, respec- [25] T.C. Moore, S.A. Newmister, I. Rayment, J.C. Escalante-Semerena, Biochemistry
51 (2012) 9647–9657.
tively.
[26] M. St. Maurice, P. Mera, K. Park, T.C. Brunold, J.C. Escalante-Semerena, I. Ray-
Cbl(I) is a potent photoreductant [209]. Photoexcitation of Cbl(I) ment, Biochemistry 47 (2008) 5755–5766.
under the action of the laser pulse ( = 532 nm) occurs at a high rate [27] M.D. Liptak, T.C. Brunold, J. Am. Chem. Soc. 128 (2006) 9144–9156.
(>108 s−1 ), after which the system returns within microseconds to [28] M. Kumar, H. Hirao, P.M. Kozlowski, J. Biol. Inorg. Chem. 17 (2012) 1107–1121.
[29] M. Kumar, P.M. Kozlowski, Angew. Chem. Int. Ed. 50 (2011) 8702–8705.
the ground state. The value of the redox potential of the excited [30] M. Kumar, N. Kumar, H. Hirao, P.M. Kozlowski, Inorg. Chem. 51 (2012)
Co(I)-state (∼–1.9 V vs. NHE) is more than 2 times greater than 5533–5538.
that of the ground state, suggesting that the photoexcited Cbl(I) [31] M. Kumar, P.M. Kozlowski, J. Inorg. Biochem. 126 (2013) 26–34.
[32] D. Lexa, J.-M. Saveant, Acc. Chem. Res. 16 (1983) 235–243.
may offer interesting chemistry with substrates to which Cbl(I) is [33] N.B. Nazhat, B.T. Golding, G.R. Alastair Johnson, P. Jones, J. Inorg. Biochem. 36
otherwise known to be inert. (1989) 75–81.
[34] D.S. Salnikov, I.A. Dereven’kov, S.V. Makarov, E.N. Artyushina, Izv. Vyssh.
Uchebn. Zaved. Khim. Khim. Tekhnol. 54 (2011) 43–46 (in Russian).
[35] I.A. Dereven’kov, D.S. Salnikov, N.I. Shpagilev, S.V. Makarov, E.V. Tarakanova,
7. Conclusion Macroheterocycles 5 (2012) 260–265.
[36] K.L. Brown, J.M. Hakimi, D.M. Nuss, Y.D. Montejano, D.W. Jacobsen, Inorg.
This review discusses the most important results of the research Chem. 23 (1984) 1463–1471.
[37] D.E. Linn Jr., E.S. Gould, Inorg. Chem. 27 (1988) 1625–1628.
on the redox reactions of cobalamin and its derivatives. We inten-
[38] F. Nome, J.H. Fendler, J. Chem. Soc., Dalton Trans. (1976) 1212–1219.
tionally did not discuss in detail here corrin modification processes [39] I. Ahmad, K. Qadeer, S. Zahid, M.A. Sheraz, T. Ismail, W. Hussain, I.A. Ansari,
in which redox reactions play a decisive role, since these processes AAPS PharmSciTech 15 (2014) 1324–1333.
[40] L. Xia, A.G. Cregan, L.A. Berben, N.E. Brasch, Inorg. Chem. 43 (2004) 6848–6857.
have been recently reviewed by Gryko and coworkers [4]. A sepa-
[41] R. Blackburn, M. Kyaw, A.J. Swallow, J. Chem. Soc., Faraday Trans. 1 (73) (1977)
rate discussion may be dedicated to the huge amount of data about 250–255.
the role of different redox states of cobalamin in food systems, bio- [42] M.N. Malinkina, A.S. Makarova, S.V. Makarov, D.S. Salnikov, Macroheterocy-
chemistry and medicine. In all these areas of redox transformations, cles 4 (2011) 42–46 (in Russian).
[43] L. Hannibal, C.A. Smith, D.W. Jacobsen, Inorg. Chem. 49 (2010) 9921–9927.
cobalamin plays a very important role. Among others, successive [44] E. Suarez-Moreira, L. Hannibal, C.A. Smith, R.A. Chavez, D.W. Jacobsen, N.E.
reduction and oxidation (a combination of the reducing capacity Brasch, Dalton Trans. (2006) 5269–5277.
with the H2 O2 -generating capacity of various nutrients) could com- [45] R.K. Suto, N.E. Brasch, O.P. Anderson, R.G. Finke, Inorg. Chem. 40 (2001)
2686–2692.
pletely account for the B12 losses in food systems [127]. While not [46] R. Mukherjee, A. McCaddon, C.A. Smith, N.E. Brasch, Inorg. Chem. 48 (2009)
providing exhaustive coverage of the subject, we nevertheless hope 9526–9534.
that this review shows the importance of studying the reactions [47] L.A. Schumacher, R. Mukherjee, J.M. Brown, H. Subedi, N.E. Brasch, Eur. J. Inorg.
Chem. (2011) 4717–4720.
leading to the formation as well as the oxidation of the reduced [48] S. Ramasamy, T.K. Kundu, W. Antholine, P.T. Manoharan, J.M. Rifkind, J. Por-
forms cobalamin, and will be useful to researchers involved in the phyrins Phthalocyanines 16 (2012) 25–38.
redox chemistry of vitamin B12 and its analogs. [49] C.B. Perry, H.M. Marques, S. Afr, J. Chem. 58 (2005) 9–15.
82 I.A. Dereven’kov et al. / Coordination Chemistry Reviews 309 (2016) 68–83

[50] H.A.O. Hill, J.M. Pratt, R.G. Thorp, B. Ward, R.J.P. Williams, Biochem. J. 120 [102] C. Wedemeyer-Exl, T. Darbre, R. Keese, Org. Biomol. Chem. 5 (2007)
(1970) 263–269. 2119–2128.
[51] K.S. Conrad, T.C. Brunold, Inorg. Chem. 50 (2011) 8755–8766. [103] J.M. Pratt, M.S.A. Hamza, G.J. Buist, J. Chem. Soc., Chem. Commun. (1993)
[52] A.S. Eisenberg, I.V. Likhtina, V.S. Znamenskiy, R.L. Birke, J. Phys. Chem. A 116 701–702.
(2012) 6851–6869. [104] S.V. Makarov, E.V. Kudrik, E.V. Naidenko, Russ. J. Inorg. Chem. 51 (2006)
[53] D.S. Salnikov, P.N. Kucherenko, I.A. Dereven’kov, S.V. Makarov, R. van Eldik, 1149–1152.
Eur. J. Inorg. Chem. (2014) 852–862. [105] H. Shimakoshi, L. Li, M. Nishi, Y. Hisaeda, Chem. Commun. 47 (2011)
[54] D.S. Salnikov, S.V. Makarov, R. van Eldik, P.N. Kucherenko, G.R. Boss, Eur. J. 10921–10923.
Inorg. Chem. (2014) 4123–4133. [106] H. Shimakoshi, M. Tokunaga, T. Baba, Y. Hisaeda, Chem. Commun. (2004)
[55] T.N. Das, R.E. Huie, P. Neta, S. Padmaja, Phys. Chem. A 103 (1999) 5221–5226. 1806–1807.
[56] M. Chemaly, Dalton Trans. (2008) 5766–5773. [107] H. Shimakoshi, M. Nishi, A. Tanaka, K. Chikama, Y. Hisaeda, Chem. Commun.
[57] K. Tahara, A. Matsuzaki, T. Masuko, J. Kikuchi, Y. Hisaeda, Dalton Trans. 42 47 (2011) 6548–6550.
(2013) 6410–6416. [108] J. Xu, H. Shimakoshi, Y. Hisaeda, J. Organomet. Chem. 782 (2015) 89–95.
[58] D.H. Truong, A. Mihajlovic, P. Gunness, W. Hindmarsh, P.J. O’Brien, Toxicology [109] K. Tahara, Y. Hisaeda, Green Chem. 13 (2011) 558–561.
242 (2007) 16–22. [110] K. Tahara, K. Mikuriya, T. Masuko, J. Kikuchi, Y. Hisaeda, J. Porphyrins Phthalo-
[59] P. Haouzi, T. Sonobe, N. Torsell-Tubbs, B. Prokopczyk, B. Chenuel, C.M. Klinger- cyanines 17 (2013) 135–141.
man, Toxicol. Sci. 141 (2014) 493–504. [111] H. Shimakoshi, M. Abiru, S. Izumi, Y. Hisaeda, Chem. Commun. (2009)
[60] P. Haouzi, B. Chenuel, T. Sonobe, Clin. Toxicol. 53 (2015) 28–36. 6427–6429.
[61] M. Brenner, S. Benavides, S.B. Mahon, J. Lee, D. Yoon, D. Mukai, M. Viseroi, A. [112] H. Shimakoshi, Y. Hisaeda, ChemPlusChem 79 (2014) 1250–1253.
Chan, J. Jiang, N. Narula, S.M. Azer, C. Alexander, G.R. Boss, Clin. Toxicol. 52 [113] K. Park, P.E. Mera, T.C. Moore, J.C. Escalante-Semerena, T.C. Brunold, Angew.
(2014) 490–497. Chem. Int. Ed. 54 (2015) 7158–7161.
[62] Y. Fujita, Y. Fujino, M. Onodera, S. Kikuchi, T. Kikkawa, Y. Inoue, H. Niitsu, K. [114] L. Brammer, D. Zhao, F.T. Ladipo, J. Braddock-Wilking, Acta Crystallogr. B51
Takahashi, S. Endo, J. Anal. Toxicol. 35 (2011) 119–123. (1995) 632–640.
[63] A. Van de Louw, P. Haouzi, Antioxid. Redox Signal. 19 (2013) 510–516. [115] W. Meister, S.E. Hennig, J.-H. Jeoung, F. Lendzian, H. Dobbek, P. Hildebrandt,
[64] P. Haouzi, B. Chenuel, T. Sonobe, C.M. Klingerman, Clin. Toxicol. 52 (2014) Biochemistry 51 (2012) 7040–7042.
566. [116] S.E. Hennig, S. Goetzl, J.-H. Jeoung, M. Bommer, F. Lendzian, P. Hildebrandt,
[65] T.-L.C. Hsu, N.E. Brasch, R.G. Finke, Inorg. Chem. 37 (1998) 5109–5116. H. Dobbek, Nat. Commun. 5 (2014), http://dx.doi.org/10.1038/ncomms5626,
[66] D.S. Salnikov, I.A. Dereven’kov, S.V. Makarov, E.S. Ageeva, A. Lupan, M. Surd- Article number 4626.
ucan, R. Silaghi-Dumitrescu, Rev. Roum. Chim. 57 (2012) 353–359. [117] T. Svetlitchnaia, V. Svetlitchnyi, O. Meyer, H. Dobbek, PNAS 103 (2006)
[67] I.A. Dereven’kov, Reactions of cobalamins and cobinamide with sulfur- 14331–14336.
containing reductants, thiocyanate and monosaccharides (Ph.D. thesis), 2013, [118] T.A. Stich, N.R. Buan, T.C. Brunold, J. Am. Chem. Soc. 126 (2004) 9735–9749.
Ivanovo, 140 pp. (in Russian). [119] J.H. Bayston, N. Kelso King, F.D. Looney, M.E. Winfield, J. Am. Chem. Soc. 91
[68] S.V. Makarov, A.K. Horvath, R. Silaghi-Dumitrescu, Q. Gao, Chem. Eur. J. 20 (1969) 2775–2779.
(2014) 14164–14176. [120] E. Jörin, A. Schweiger, Hs.H. Günthard, J. Am. Chem. Soc. 105 (1983)
[69] S.V. Makarov, R. Silaghi-Dumitrescu, J. Sulfur Chem. 34 (2013) 444–449. 4277–4286.
[70] D. Lexa, J.M. Saveant, J. Zickler, J. Am. Chem. Soc. 102 (1980) 2654–2663. [121] A. von Zelewsky, Helv. Chim. Acta 55 (1972) 2941–2947.
[71] R.L. Birke, Q. Huang, T. Spataru, D.K. Gosser Jr., J. Am. Chem. Soc. 128 (2006) [122] S. Van Doorslaer, A. Schweiger, B. Kräutler, J. Phys. Chem. B 105 (2001)
1922–1936. 7554–7563.
[72] M. Kumar, W. Galezowski, P.M. Kozlowski, Int. J. Quantum Chem. 113 (2013) [123] M. Surducan, S.V. Makarov, R. Silaghi-Dumitrescu, Polyhedron 78 (2014)
479–488. 72–84.
[73] D.S. Salnikov, I.A. Dereven’kov, E.N. Artyushina, S.V. Makarov, Russ. J. Phys. [124] E. Hohenester, C. Kratky, B. Kräutler, J. Am. Chem. Soc. 113 (1991) 4523–4530.
Chem. A 87 (2013) 52–56. [125] E. Suarez-Moreira, J. Yun, C.S. Birch, J.H.H. Williams, A. McCaddon, N.E. Brasch,
[74] M.S.A. Hamza, X. Zou, K.L. Brown, R. van Eldik, J. Chem. Soc., Dalton Trans. J. Am. Chem. Soc. 131 (2009) 15078–15079.
(2002) 3832–3839. [126] R.S. Dassanayake, D.E. Cabelli, N.E. Brasch, ChemBioChem 14 (2013)
[75] J. Kim, C. Gherasim, R. Banerjee, PNAS 105 (2008) 14551–14554. 1081–1083.
[76] Z. Li, C. Gherasim, N.A. Lesniak, R. Banerjee, J. Biol. Chem. 289 (2014) [127] P.W. Johns, A. Das, E.M. Kuil, W.A. Jacobs, K.J. Schimpf, D.J. Schmitz, Int. J. Food
16487–16497. Sci. Technol. 50 (2015) 421–430.
[77] D.S. Froese, J. Zhang, S. Healy, R.A. Gravel, Mol. Genet. Metab. 98 (2009) [128] A. Nicolaou, S.H. Kenyon, J.M. Gibbons, T. Ast, W.A. Gibbons, Eur. J. Clin. Invest.
338–343. 26 (1996) 167–170.
[78] J. Jeong, J. Kim, Biochem. Biophys. Res. Commun. 412 (2011) 360–365. [129] A. Kambo, V.S. Sharma, D.E. Casteel, V.L. Woods Jr., R.B. Pilz, G.R. Boss, J. Biol.
[79] G.N. Schrauzer, J.A. Seck, T.M. Beckham, Bioinorg. Chem. 2 (1973) 211–229. Chem. 280 (2005) 10073–10082.
[80] D. Lexa, J.-M. Saveant, J. Am. Chem. Soc. 100 (1978) 3220–3222. [130] M. Brouwer, W. Chamulitrat, G. Ferruzzi, D.L. Sauls, J.B. Weinberg, Blood 88
[81] H.P.C. Hogenkamp, G.T. Bratt, A.T. Kotchevar, Biochemistry 26 (1987) (1996) 1857–1864.
4123–4127. [131] R. Schubert, U. Krien, I. Wulfsen, D. Schiemann, G. Lehmann, N. Ulfig, R.W.
[82] D.W. Jacobsen, E.H. Pezacka, K.L. Brown, J. Inorg. Biochem. 50 (1993) 47–63. Veh, J.R. Schwarz, H. Gago, Hypertension 43 (2004) 891–896.
[83] H. Shimakoshi, M. Tokunaga, K. Kuroiwa, N. Kimizuka, Y. Hisaeda, Chem. [132] F. Jiang, C.G. Li, M.J. Rand, Eur. J. Pharmacol. 340 (1997) 181–186.
Commun. (2004) 50–51. [133] M. Wolak, A. Zahl, T. Schneppensieper, G. Stochel, R. van Eldik, J. Am. Chem.
[84] A.J. Brooks, M. Vlasie, R. Banerjee, T.C. Brunold, J. Am. Chem. Soc. 126 (2004) Soc. 123 (2001) 9780–9791.
8167–8180. [134] D. Zheng, R.L. Birke, J. Am. Chem. Soc. 123 (2001) 4637–4638.
[85] T.A. Stich, A.J. Brooks, N.R. Buan, T.C. Brunold, J. Am. Chem. Soc. 125 (2003) [135] L. Hannibal, C.A. Smith, D.W. Jacobsen, N.E. Brasch, Angew. Chem. Int. Ed. 46
5897–5914. (2007) 5140–5143.
[86] L. Hannibal, J. Kim, N.E. Brasch, S. Wange, D.S. Rosenblatt, R. Banerjee, D.W. [136] H.A. Hassanin, M.F. El-Shahat, S. DeBeer, C.A. Smith, N.E. Brasch, Dalton Trans.
Jacobsen, Mol. Genet. Metab. 97 (2009) 260–266. 39 (2010) 10626–10630.
[87] J. Kim, L. Hannibal, C. Gherasim, D.W. Jacobsen, R. Banerjee, J. Biol. Chem. 284 [137] I.G. Pallares, T.C. Brunold, Inorg. Chem. 53 (2014) 7676–7691.
(2009) 33418–33424. [138] V.S. Sharma, R.B. Pilz, G.R. Boss, D. Magde, Biochemistry 42 (2003) 8900–8908.
[88] Z. Li, N.A. Lesniak, R. Banerjee, J. Am. Chem. Soc. 136 (2014) 16108–16111. [139] K.E. Broderick, L. Alvarez, M. Balasubramanian, D.D. Belke, A. Makino, A. Chan,
[89] M. Ruetz, C. Gherasim, K. Gruber, S. Fedosov, R. Banerjee, B. Kräutler, Angew. V.L. Woods Jr., W.H. Dillmann, V.S. Sharma, R.B. Pilz, T.D. Bigby, G.R. Boss, Exp.
Chem. Int. Ed. 52 (2013) 2606–2610. Biol. Med. (Maywood, NJ) 232 (2007) 1432–1440.
[90] M. Ruetz, R. Salchner, K. Wurst, S. Fedosov, B. Kräutler, Angew. Chem. Int. Ed. [140] H. Subedi, N.E. Brasch, Inorg. Chem. 52 (2013) 11608–11617.
52 (2013) 11406–11409. [141] R. Spitler, R. Schwappacher, T. Wuc, X. Kong, K. Yokomori, R.B. Pilz, G.R. Boss,
[91] P.M. Kozlowski, J. Kuta, W. Galezowski, J. Phys. Chem. B 111 (2007) M.W. Berns, Cell. Signal. 25 (2013) 2374–2382.
7638–7645. [142] K.E. Broderick, V. Singh, S. Zhuang, A. Kambo, J.C. Chen, V.S. Sharma, R.B. Pilz,
[92] B.D. Martin, R.G. Finke, J. Am. Chem. Soc. 114 (1992) 585–592. G.R. Boss, J. Biol. Chem. 280 (2005) 8678–8685.
[93] J.E. Argüello, C. Costentin, S. Griveau, J.-M. Saveant, J. Am. Chem. Soc. 127 [143] D. Zheng, R.L. Birke, J. Am. Chem. Soc. 124 (2002) 9066–9067.
(2005) 5049–5055. [144] M. Wolak, G. Stochel, R. van Eldik, Inorg. Chem. 45 (2006) 1367–1379.
[94] H. Shimakoshi, M. Tokunaga, Y. Hisaeda, Dalton Trans. (2004) 878–882. [145] H.A. Hassanin, L. Hannibal, D.W. Jacobsen, M.F. El-Shahat, M.S.A. Hamza, N.E.
[95] L. Pan, H. Shimakoshi, T. Masuko, Y. Hisaeda, Dalton Trans. (2009) 9898–9905. Brasch, Angew. Chem. Int. Ed. 48 (2009) 8909–8913.
[96] K.P. Jensen, U. Ryde, J. Am. Chem. Soc. 125 (2003) 13970–13971. [146] H. Subedi, H.A. Hassanin, N.E. Brasch, Inorg. Chem. 53 (2014) 1570–1577.
[97] M. Alfonso-Prieto, X. Biarnes, M. Kumar, C. Rovira, P.M. Kozlowski, J. Phys. [147] F. Roncaroli, T.E. Shubina, T. Clark, R. van Eldik, Inorg. Chem. 45 (2006)
Chem. B 114 (2010) 12965–12971. 7869–7876.
[98] H. Olteanu, R. Banerjee, J. Biol. Chem. 276 (2001) 35558–35563. [148] M. Wolak, G. Stochel, M. Hamza, R. van Eldik, Inorg. Chem. 39 (2000)
[99] H. Olteanu, K.R. Wolthers, A.W. Munro, N.S. Scrutton, R. Banerjee, Biochem- 2018–2019.
istry 43 (2004) 1988–1997. [149] M. Wolak, G. Stochel, R. van Eldik, J. Am. Chem. Soc. 125 (2003) 1334–1351.
[100] D. Lexa, J.M. Saveant, J. Am. Chem. Soc. 98 (1976) 2652–2658. [150] R.S. Dassanayake, D.E. Cabelli, N.E. Brasch, J. Inorg. Biochem. 142 (2015) 54–56.
[101] J.S. Dorweiler, R.G. Finke, R.G. Matthews, Biochemistry 42 (2003) [151] R. Mukherjee, N.E. Brasch, Chem. Eur. J. 17 (2011) 11805–11812.
14653–14662. [152] H. Subedi, N.E. Brasch, Eur. J. Inorg. Chem. (2015) 3825–3834.
I.A. Dereven’kov et al. / Coordination Chemistry Reviews 309 (2016) 68–83 83

[153] R.S. Dassanayake, J.T. Shelley, D.E. Cabelli, N.E. Brasch, Chem. Eur. J. 21 (2015) [183] T.C. Tallant, L. Paul, J.A. Krzycki, J. Biol. Chem. 276 (2001) 4485–4493.
6409–6419. [184] D. Naidu, S.W. Ragsdale, J. Bacteriol. 183 (2001) 3276–3281.
[154] C.S. Birch, N.E. Brasch, A. McCaddon, J.H.H. Williams, Free Radic. Biol. Med. 47 [185] J. Haglund, A. Rafiq, L. Ehrenberg, B.T. Golding, Chem. Res. Toxicol. 13 (2000)
(2009) 184–188. 253–256.
[155] I.A. Dereven’kov, L. Hannibal, D.S. Salnikov, T.T. Bui Thi, S.V. Makarov, I. [186] J.M. Pratt, P.R. Norris, M.S.A. Hamza, R. Bolton, J. Chem. Soc. Chem. Commun.
Ivanović-Burmazović, submitted to Angew. Chem. Int. Ed. (2015). (1994) 1333–1334.
[156] I.A. Dereven’kov, T.T. Bui Thi, D.S. Salnikov, S.V. Makarov, accepted to Russ. J. [187] R. Banerjee, S.W. Ragsdale, Annu. Rev. Biochem. 72 (2003) 209–247.
Phys. Chem. A (2016), doi:10.7868/S0044453716030080. [188] R.G. Matthews, Acc. Chem. Res. 34 (2001) 681–689.
[157] S. Cockle, H.A.O. Hill, S. Ridsdale, R.J.P. Williams, J. Chem. Soc., Dalton Trans. [189] N. Kumar, P.M. Kozlowski, J. Phys. Chem. B 117 (2013) 16044–16057.
(1972) 297–302. [190] S.-L. Chen, M.R.A. Blomberg, P.E.M. Siegbahn, J. Phys. Chem. B 115 (2011)
[158] J.H. Bayston, F.D. Looney, J.R. Pilbrow, M.E. Winfield, Biochemistry 9 (1970) 4066–4077.
2164–2172. [191] H.V. Motwani, M. Törnqvist, J. Chromatogr. A 1218 (2011) 4389–4394.
[159] P.N. Balasubramanian, E.S. Gould, Inorg. Chem. 24 (1985) 1791–1793. [192] J. Haglund, V. Silvari, E. Esmans, M. Törnqvist, J. Chromatogr. A 1119 (2006)
[160] R.G. Pearson, H. Sobel, J. Songstad, J. Am. Chem. Soc. 90 (1968) 319–326. 246–250.
[161] G.N. Schrauzer, E. Deutsch, R.J. Windgassen, J. Am. Chem. Soc. 90 (1968) [193] J. Haglund, A.-L. Magnusson, L. Ehrenberg, M. Törnqvist, Toxicol. Environ.
2441–2442. Chem. 85 (2003) 81–94.
[162] K.P. Jensen, J. Phys. Chem. B 109 (2005) 10505–10512. [194] J. Shey, W.A. van der Donk, J. Am. Chem. Soc. 122 (2000) 12403–12404.
[163] N. Kumar, M. Alfonso-Prieto, C. Rovira, P. Lodowski, M. Jaworska, P.M. [195] K.M. McCauley, S.R. Wilson, W.A. van der Donk, Inorg. Chem. 41 (2002)
Kozlowski, J. Chem. Theor. Comput. 7 (2011) 1541–1551. 5844–5848.
[164] S.M. Chemaly, R.A. Hasty, J.M. Pratt, J. Chem. Soc. Dalton Trans. (1983) [196] K.M. McCauley, S.R. Wilson, W.A. van der Donk, J. Am. Chem. Soc. 125 (2003)
2223–2227. 4410–4411.
[165] D. Lexa, J.-M. Saveant, J. Chem. Soc. Chem. Commun. (1975) 872–874. [197] D.A. Pratt, W.A. van der Donk, Chem. Commun. (2006) 558–560.
[166] W.D. Robertson, A.M. Bovell, K. Warncke, J. Biol. Inorg. Chem. 18 (2013) [198] S. Kliegman, K. McNeill, Environ. Sci. Technol. 43 (2009) 8961–8967.
701–713. [199] K. Tahara, H. Shimakoshi, A. Tanaka, Y. Hisaeda, Bull. Chem. Soc. Jpn. 83 (2010)
[167] P.N. Balasubramanian, E.S. Gould, Inorg. Chem. 22 (1983) 2635–2637. 1439–1446.
[168] N.T. Plymale, R.S. Dassanayake, H.A. Hassanin, N.E. Brasch, Eur. J. Inorg. Chem. [200] Md.A. Jabbar, H. Shimakoshi, Y. Hisaeda, Chem. Commun. (2007) 1653–1655.
(2012) 913–921. [201] H.V. Motwani, S. Qiu, B.T. Golding, H. Kylin, M. Törnqvist, Food Chem. Toxicol.
[169] R. Mukherjee, N.E. Brasch, Chem. Eur. J. 17 (2011) 11723–11727. 49 (2011) 750–757.
[170] D. Zheng, L. Yan, R.L. Birke, Inorg. Chem. 41 (2002) 2548–2555. [202] H.V. Motwani, H. Shimakoshi, B.T. Golding, M. Törnqvist, Y. Hisaeda, Tetrahe-
[171] S.L. Vilakazi, T. Nyokong, Electrochim. Acta 46 (2000) 453–461. dron Lett. 55 (2014) 2667–2670.
[172] R. Silaghi-Dumitrescu, Inorg. Chem. 43 (2004) 3715–3718. [203] J. Shey, C.M. McGinley, K.M. McCauley, A.S. Dearth, B.T. Young, W.A. van der
[173] N. Xu, J. Yi, G.B. Richter-Addo, Inorg. Chem. 49 (2010) 6253–6266. Donk, J. Org. Chem. 67 (2002) 837–846.
[174] R.G.S. Banks, R.J. Henderson, J.M. Pratt, J. Chem. Soc. A (1968) 2886–2889. [204] S. Kliegman, K. McNeill, Dalton Trans. (2008) 4191–4201.
[175] P.N. Balasubramanian, E.S. Gould, Inorg. Chem. 23 (1984) 824–828. [205] S. Cretnik, K.A. Thoreson, A. Bernstein, K. Ebert, D. Buchner, C. Laskov, S.
[176] I.A. Dereven’kov, D.S. Salnikov, S.V. Makarov, G.R. Boss, O.I. Koifman, Dalton Haderlein, O. Shouakar-Stash, S. Kliegman, K. McNeill, M. Elsner, Environ. Sci.
Trans. 42 (2013) 15307–15316. Technol. 47 (2013) 6855–6863.
[177] G.C. Pillai, E.S. Gould, Inorg. Chem. 25 (1986) 3353–3356. [206] M. Bommer, C. Kunze, J. Fesseler, T. Schubert, G. Diekert, H. Dobbek, Science
[178] G.C. Pillai, R.N. Bose, E.S. Gould, Inorg. Chem. 26 (1987) 3120–3123. 346 (2014) 455–458.
[179] R.V. Banerjee, V. Frasca, D.P. Ballou, R.G. Matthews, Biochemistry 29 (1990) [207] A. Parthasarathy, T.A. Stich, S.T. Lohner, A. Lesnefsky, R.D. Britt, A.M.
11101–11109. Spormann, J. Am. Chem. Soc. 137 (2015) 3525–3532.
[180] S.A. Burke, S.L. Lo, J.A. Krzycki, J. Bacteriol. 180 (1998) 3432–3440. [208] K.A.P. Payne, C.P. Quezada, K. Fisher, M.S. Dunstan, F.A. Collins, H. Sjuts, C.
[181] D.J. Ferguson Jr., N. Gorlatova, D.A. Grahame, J.A. Krzycki, J. Biol. Chem. 275 Levy, S. Hay, S.E.J. Rigby, D. Leys, Nature 517 (2015) 513–516.
(2000) 29053–29060. [209] D. Achey, E.C. Brigham, B.N. DiMarco, G.J. Meyer, Chem. Commun. 50 (2014)
[182] L. Paul, D.J. Ferguson, J.A. Kryzycki, J. Bacteriol. 182 (2000) 2520–2529. 13304–13306.

You might also like