The Chemistry and Biochemistry of Vanadium and The Biological Activities Exerted by Vanadium Compounds PDF
The Chemistry and Biochemistry of Vanadium and The Biological Activities Exerted by Vanadium Compounds PDF
849
Contents
1. Introduction
2. Aqueous V(V) Chemistry and the
PhosphateVanadate Analogy
2.1. Aqueous V(V) Chemistry
2.2. Mimicking Cellular Metabolites:
VanadatePhosphate Analogy
(Four-Coordinate Vanadium)
2.3. Structural Model Studies of Vanadate Esters
2.4. Vanadate Esters: Functional Analogues of
Phosphate Esters
2.5. Vanadate Anhydrides: Structural Analogy
with Condensed Phosphates
2.6. Potential Future Applications of
Vanadium-Containing Ground State
Analogues in Enzymology
3. Haloperoxidases: V(V) Containing Enzymes and
Modeling Studies
3.1. Bromoperoxidase
3.2. Chloroperoxidase
3.3. From Haloperoxidases to Phosphatases
3.4. Structural Modeling Studies of the
Vanadium-Dependent Haloperoxidases
3.5. Spectroscopic Modeling Studies of the
Vanadium-Dependent Haloperoxidases
3.6. Functional Modeling of the
Vanadium-Dependent Haloperoxidases
3.7. Oxidation of Sulfides by Haloperoxidases
(VHPOs)
4. Inhibition of Phosphorylases by V-Compounds:
Phosphatases, Ribonuclease, Other
Phosphorylases, and ATPases
4.1. Phosphatases
4.1.1. Vanadate: An Imperfect Transition State
Analogue of Phosphatases
4.1.2. V(IV) Chemistry and Inhibition of
Phosphatases by V(IV)
4.1.3. Inhibition of Alkaline and Acid
Phosphatases by Vanadium Compounds
4.1.4. Inhibition of Protein Phosphatases by
Vanadium Compounds
4.1.5. Inhibition of Purple Acid Phosphatases by
Vanadium Compounds
4.1.6. Phosphatase and Other Phosphorylase
Inhibition by Oxometalates
4.1.7. Structural Models of the Transition State
of Phosphate Ester Hydrolysis
850
851
851
852
852
853
854
855
855
855
856
857
857
859
860
860
861
861
861
862
864
864
865
866
866
* To whom co rrespondence should be addressed. Phone: 970-4917635. Fax: 970-491-1801. E-mail: [email protected].
4.2. Ribonuclease
4.2.1. Structural Characterization of Model
Compounds for Inhibitors of Ribonuclease
4.2.2. Characterization of the
Nucleoside-Vanadate Complexes that
Form in Solution and Inhibit Ribonuclease
4.2.3. VanadiumNucleoside Complexes:
Functional Inhibitors of Ribonuclease
4.3. Other Phosphorylases
4.4. ATPases
4.4.1. Structural Precedence for the
VanadatePhosphate Anhydride Unit:
Five- or Six-Coordinate Vanadium
4.4.2. Vanadate as a Photocleavage Agent for
ATPases
5. Amavadine and Siderophores
5.1. Amavadine
5.1.1. Amavadine: Structure
5.1.2. Amavadine: Activities and Roles
5.2. Siderophores
5.2.1. Effects of Vanadium on
Siderophore-Mediated Iron Transport
5.2.2. Characterization of
VanadiumSiderophore Complexes and
Model Complexes
5.2.3. Vanadium Citrate Complexes: Structure
and Speciation
5.2.4. V(V) and V(IV) Hydroxamate Complexes
6. Tunicates and the Polychaete Fan Worm
6.1. Tunicates
6.1.1. Location of Vanadium in Tunicate Blood
Cells
6.1.2. Aqueous V(III) Chemistry
6.1.3. Oxidation State of Vanadium in Tunicates
6.1.4. Uptake of Vanadate into Tunicates
6.1.5. Vanadium Binding Proteins: Vanabins
6.1.6. Model Complexes and Their Chemistry
6.1.7. Catechol-Based Model Chemistry
6.1.8. Vanadium Sulfate Complexes
6.2. Fan Worm Pseudopotamilla occelata
7. Vanadium Nitrogenase
7.1. Nitrogenases
7.2. Biochemistry of Nitrogenase
7.3. Clusters in Nitrogenase and Model Systems:
Structure and Reactivity
7.4. Structural Model Complexes of the VFe
Cofactor Binding Site
7.5. Homocitrate
7.6. Activation of Nitrogen
866
867
867
868
868
869
869
870
871
871
871
872
873
873
874
874
875
876
876
877
877
878
879
879
880
880
881
883
883
883
884
885
886
887
887
Crans et al.
888
888
888
889
889
889
891
892
892
892
892
893
1. Introduction
Vanadium is a trace element, which may be beneficial and possibly essential in humans1 but certainly
essential for some living organisms.2-11 Metal ions
and thus vanadium ions can play a role in biology as
counterions for protein, DNA, RNA, and in various
biological organelles. The structural role is often
manifested by the maintenance of various biological
structures, whereas a functional role is to bring key
reactivity to a reaction center for a protein. Vanadium ions have many structural roles reflected by its
structural and electronic analogy to phosphorus.9,12-20
In addition, the vanadium ion is an enzyme cofactor,7,9,21-30 and is found in certain tunicates4-11,16
and possibly mammals.1 Reviews on how vanadium
can act and function in the biosphere include investigations into the fundamental coordination and
redox chemistry of the element,16-18,29,31-35 as well as
structural and functional aspects of biological systems and/or metabolites.12,36 Modeling biological activities of various types have long been of interest to
chemists, with this discipline focusing on the structural modeling until about a decade ago when the
focus shifted to functional modeling. Clearly modeling
that includes both aspects will be most informative,
and the ultimate goals for model chemists. Although
the latter in general may be of greater interest at
the present time, the structural aspects of the various
oxidation states are defining its effects in many
biological systems. In this review, we have combined
the two fundamentally different aspects of modeling
because the coverage of only one of these areas in
our opinion would not provide the reader the proper
sense of the effects and activities exerted by vanadium compounds (V-compounds).
This review describes the voyage from the discovery of the first vanadium-containing enzymes in
1984, haloperoxidases,7,9,21-26,37,38 to the current X-ray
crystallographic studies; it is a fascinating story that
provides inspiration to chemist in many fields. These
studies map out the enzyme active site24,39-43 and
demonstrate the structural and functional link between apohaloperoxidases and certain phosphatases.24,39-43 These discoveries now give the bioinorganic chemists clear directives. The detailed mecha-
the second V-containing enzyme, the V-nitrogenase,7,27,28,51 is not as well understood as the vanadiumcontaining haloperoxidases, it is currently a target
for both mechanistic7,27-29 and modeling52,53 studies
with novel structural information constantly modifying the current understanding this system. Combined, these studies bode well for new classes of
V-containing proteins, the vanabins,2,5,10,11 which
have been proposed to be transport proteins for
vanadium, and promise exciting developments in the
future.
The elucidation of the structure of the vanadiumcontaining natural product, amavadine,54 remains an
interesting structural investigation,29,55-58 although
questions now being addressed in this area have
shifted to the investigations of this complexs catalytic
properties. Perhaps these developments will provide
additional clues to the function of this unique compound. The investigations into tunicate metabolism
and biology,2,3,11,59,60 the recognition of the important
role of sulfate in the intact blood cells, and the firm
documentation that V(III) does exist in some species
as well as the discovery of a V-containing transport
protein present key developments to understanding
why and how these organisms contain vanadium.
Finally, the boom of investigations into the insulinlike action of V-complexes documents the ability of a
variety of V-complexes to lower elevated glucose
levels in diabetic animals treated with V-compounds.30,61-64 Although only V(IV)-compounds are
generally considered in this regard, reports of vanadium in oxidation states V and III are likely to firmly
establish the general insulin-like activity of V-compounds.30,63-68 While insulin-like activity of V-compounds is linked to the inhibitory effects of Vcompounds of phosphatases, undoubtedly other activities will emerge now that the vanadium-dependent
haloperoxidase and phosphatase association has been
made. Indeed, we summarize briefly other enzyme
activities that have been reported with V-compounds
in the final section of the review. The many discoveries reviewed here document the versatile nature of
(3)
Around neutral pH, H2VO4- and HVO42- oligomerize to form dimeric (4), tetrameric (5), and pentameric
Crans et al.
(4)
ROPO3H- + H2VO4- h
ROP(O)2OV(O)2OH2- + H2O (5)
Crans et al.
(6)
(7)
3.1. Bromoperoxidase
Although peroxidase activity has been known in
marine extracts for many years,26 it was not until
1984 when Vilter discovered that vanadium content
was required in a bromoperoxidase from Ascophyllum
nodosum.25,130 The enzyme mechanism was examined
using structural, kinetic, and spectroscopic methods,
and issues such as the oxidation state of the metal
and its coordination geometry have been investigated
in detail. The oxidation potentials of the halides are
pH-dependent, and, in general, more acidic pH values
are required to oxidize the more electronegative
halides; this suggests that there is some flexibility
in these enzyme systems. Substrate specificity and
product distribution remain key areas of interest, and
Crans et al.
3.2. Chloroperoxidase
Haloperoxidases have been isolated from red, brown,
and green algae,7,26 seaweed,7,26 a lichen,7,147 and a
VCPO from terrestrial fungi.47 In the first X-ray
structure of VCPO from the fungus C. inaequalis, the
vanadium is bound as monoprotonated vanadate(V)
in a trigonal-bipyramidal coordination geometry.134
The metal is coordinated to three oxygens in the
equatorial plane, to an OH group at one apical
position, and to the N nitrogen of a histidine at the
other apical position. The key residues in the active
site are Lys353, Arg360, His404, Arg490, and His496
(Figure 2, bottom). The protein fold is mainly R-helical with two four-helix bundles as the main structural
motifs. An amino acid sequence comparison with
VBPO from the seaweed As. nodosum shows high
similarities in the metal binding site. All residues
interacting with vanadate(V) are conserved except for
Lys353, which is an asparagine in VBPO.134,148 The
crystal structure has been determined for the apoenzyme,135,149 and with vanadate,134 tungstate,150 and
peroxovanadate135,149 in the active site. Although
overall the two structures of the vanadium- and
tungsten-bound proteins are virtually identical, one
major difference exists in the active site. In the
vanadium-protein structure, the vanadium-histidine bond is strong, whereas in the tungstateprotein bound structure such bonding is either lacking or very weak. This observation demonstrates the
fine-tuning that exists in this system for metalcofactor binding.150 Investigations into the V-binding
in native and mutant chloroperoxidase show that the
latter holds the vanadium ion less tightly; however,
the mutant maintains bromoperoxidase activity.151
The X-ray structures of the chloroperoxidase from C.
inaequalis and active site mutants document the
intricate H-bonding network that is in place to
Crans et al.
using electrospray ionization-mass spectrometry (ESIMS), 51V NMR spectroscopy, and ab initio calculations has been found to be an excellent tool for
obtaining direct information about the structure and
chemistry of peroxovanadates in solution.44,231,232
Table 1. Summary of the Structural Features of the Crystallographically Characterized Peroxo Model
Complexesa,b,c
ligand set
on Va
peroxo
ligands
nonperoxo,
nonoxo ligand(s)b
CN
N (NO5)
NO2 (NO5)
NO (NO6)
2
1
2
6
7d
7
NO3 (NO6)
NO3 (NO6)
NO3 (NO6)
N2O (N2O4)
N2 (N2O5)
N2 (N2O5)
N2O2 (N2O5)
N2O2 (N2O5)
1
1
1
1
2
2
1
1
7
7
7
6
7
7
7
N2O2 (N2O5)
N2O2 (N2O5)
1
1
7
7
N2O2 (N2O5)
N2O (N3O4)
N3O (N3O4)
1
1
7
7
N3O (N3O4)
N3O (N3O4)
N3O (N3O4)
N3O (N3O4)
N4 (N4O3)
O (O6)
O2 (O6)
1
1
1
1
1
4
2
7
7
7
7
7
6,6
6,6
O2,3 (O6,7)
7,7 or 6,6f
O (O7)
O (O7)
O2 (O7)
O2 (O7)
O4 (O7)
O5 (O7)
3
4
2
2
1
2
7
7,7
7
7
7
7,7
ammonia/imidazole
iminodiacetic acid
picolinic acid/3-hydroxy-picolinic acid/
2,4-pyridine dicarboxylic acid/
3-acetatoxy-picolinic acid
picolinic acid, H2O 2
dipicolinic acid, H2O
N-(2-hydroxyethyl)-iminodiacetic acid
glycylglycine
bipyridine (4 structures)
5-nitro-phenanthroline
phenanthroline, H2O 2
picolinic acid 2/
pyrazine-2-carboxylic acid 2
D,L-N-carboxymethyl-histidine
ethylenediamine-tetracetic acid
(two structures)
N-(carbamoylmethyl)-iminodiacetatic acid
(2 structures)/N-(carbamoylethyl)iminodiacetatic acid
1-(2-pyridylazo-)-2-naphthol, pyridine
picolinic acid, bipyridine/picolinic acid,
phenanthroline
Tris(3,5-diisopropyl-pyrazol-1-yl)borate,H2O
nitrilotriacetic acid (5 structures)
N,N-bis(2-pyridylmethyl)-glycine
N,N-bis(2-pyridylmethyl)--alanine
bipyridine 2/phenanthroline 2
H2O (three structures)
glycolic acid 2/D-lactic acid, L-lactic acid
(two structures)/L-lactic acid 2/
mandelic acid 2
citric acid (two structures)/
malic acid (four structures)
hydroxide
oxide (two structures)/hydroxide
oxalic acid (three structures)
carbonate
oxalic acid 2
L-tartaric acid 2, H2O (L)
ligand
denticity
structural
illustrationc
ref(s)
1
3
2
33
34
35
182, 183
184
185, 186
2,1,1
3,1
4
3
2
2
2,1,1
2,2
36
37
38
39
35
35
36
40
187
188
189
190
191-194
195
196
196, 197
4
4
38
38
198
199
38
200-202
3,1
2,2
37
40
203
196,204
3,1
4
4
4
2,2
1
2,2
37
38
38
38
40
41
42
205
206-209
189
143
210
211-213
214-217
3,3
43
218-221
1
1
2
2
2,2
2,2,1
44
45
35
35
40
46
222
223-225
191, 222, 226
227
228
229
a The complexes are sorted by the overall donor ligand set, which is given in parentheses after the donor ligand set of the
nonperoxo and nonoxo ligands. b The ligand sets of each structure are separated by a slash. If two (or three) of the same ligands
are present in a given structure, then the number of such ligands is denoted by 2 or 3; the number of structures of the same
complex appears in parentheses. c The key to the structural illustrations is shown in Figure 3. d The complex forms an extended
array with an adjacent VdO group serving as the seventh ligand. f The terminal carboxylate oxygens trans to the doubly bonded
oxo ligands weakly interact with the vanadium (V-Ocarboxylate distance 2.5 ).
Figure 3. Ball and stick representation of [VO(O2)(BrNH2pyg2)]-. All protons have been omitted for clarity.
Reproduced from the spatial coordinates given in ref 235.
Copyright 2002 American Chemical Society.
demonstrated, such trends do not hold when noninnocent ligands with low energy, ligand-to-metal
charge-transfer bands are used in the complexes.246
For example, catecholate and hydroxamate V(V)complexes give rise to chemical shifts that are
unusually far downfield. However, for complexes with
O and N donors such as water, hydroxide, alcohols,
monodentate carboxylates, and amines, shifts in the
range of -400 to -600 ppm are observed. Shifts
upfield of -600 ppm tend to result from negatively
charged multidentate ligands that form three- or
four-membered chelate rings such as peroxides.7,239,240
Early 51V NMR studies on the VBPO from As.
nodosum show an unusually broad signal at about
-1200 ppm and remains an observation that needs
further investigation.247 It is known that association
of the quadrupolar vanadium nucleus with a large
protein can broaden the signal width if the tumbling
of the complex becomes sufficiently slow. Considering
the current information of model studies and the
crystal structure of the various V-complexes, a protein complex is likely to contain at least three O
donors and one N (histidine) donor. Although noninnocent ligand complexes may give rise to a complex
with a -1200 ppm chemical shift, the ligands found
so far in the crystal structures of protein-V-complexes are not consistent with the reported spectrum.
Pecoraro and co-workers have spearheaded studies
using ESEEM on model complexes to explore the
coordination environment in the protein system.7
Recent ESEEM spectra of model complexes suggested
that the ESEEM of the reduced enzyme7 was consistent with the presence of one imidazole ligand in
the axial position as well as a second imidazole in
the equatorial plane of the vanadyl ion observed in
the original ESEEM spectra.248 The possibility that
V(V) binds to both histidine residues in the active
site would be consistent with the inactivity of this
form of the enzyme because the second histidine that
functions as an acid-base catalyst is firmly bound
to the protein.249 ESEEM is also useful for investigation of histidine-V-complexes modeling the histidine-V-complex in proteins.7,169 Many of the wellcharacterized model systems are Schiff bases; the
corresponding V(V) catechol complex with the ligand
HSALIMH is shown (25).250 Recently, Butler and coworkers reported the reactivity of recombinant and
mutant vanadium bromoperoxidase from Co. officinalis. Mutation of the conserved histidine residue to
an alanine (H480A) resulted in the loss of the ability
to efficiently oxidize bromide, although the ability to
oxidize iodide is retained.249
Crans et al.
reaction, complexing ligands should be able to enhance and/or retard this reaction. Since then, a range
of systems have been reported competent to carry out
this reaction.7,9,29,44-48,170,252 These complexes include
the first monomeric complexes 26 and 27 reported
by Butler and co-workers,252 which showed that the
bromination of the organic substrates proceeded
exclusively via an electrophilic mechanism and involved no radical intermediates. Other complexes
found to be competent in oxidizing bromide include
complexes from additional Schiff base ligands, nitrilotriphosphoric acid and citrate.189 Interestingly,
pyridine-2,6-dicarboxylic acid is found to protect H2O2
against reduction by bromide. Other ligands such as
picolinic acid do not form sufficiently stable adducts
to exist in the presence of H2O2.189
Pecoraro and co-workers designed a model system
with ligands which completed the coordination sphere
of an oxoperoxovanadium unit (28-31).189,253,254 The
most efficient complex of this class, [VO(O2)Hheida](28), showed some properties to be different than
those in the Butler systems; acid was required for
both catalytic and stoichiometric catalytic activity.189,253 However, in the presence of excess acid and
peroxide, up to 10 turnovers were accomplished
within 3 min, and thus increased the rate of reaction
by at least an order of magnitude greater than that
observed for previous systems. Furthermore, this
model system was the first vanadium model compound to demonstrate both the halogenation and
catalase reaction catalyzed by VBPO.254 Detailed
mechanistic studies were carried out with these
complexes, and supported a mechanism with a protonation preequilibrium preceding the halide oxidation step.189 Studies with H218O2 showed persuasively
that the oxidation of bromide produced O2, which was
completely labeled with 18O and thus indicated that
peroxide is oxidized without oxygen-oxygen bond
cleavage.189
Pecoraro and co-workers proposed a variation on
the enzyme mechansim in which an L group replaces
the EnzO3 for the catalysis of the oxidation by model
compounds.7 In this mechanism, the halide is oxidized by peroxovanadium complexes via nucleophilic
attack by the halide on a protonated oxoperoxovanadium species. The trihalide is the only product
observed under conditions of excess halide and is not
directly specified in such a mechanism recognizing
that the initial formation of hypohalous acid (HOX)
is formally equivalent to OH- + X+. This variant
mechanism is also consistent with the data reported
with Butler and co-workers when considering that
the proton-independent oxidation reaction in reality
results in the liberation of one equivalent of acid
when peroxide binds.7 A two-phase system was
designed,255 and ab initio calculations were carried
out to provide evidence for the hypobromite-like
V-complex.256
4. Inhibition of Phosphorylases by
V-Compounds: Phosphatases, Ribonuclease,
Other Phosphorylases, and ATPases
4.1. Phosphatases
Phosphatases catalyze the hydrolysis of phosphate
ester bonds (eq 8), and the mechanism involves
formation of five-coordinate, high-energy intermediates. These enzymes are generally potently inhibited
by vanadate, which is recognized to be a transition
state analogue for the phosphatase-catalyzed reaction. The inhibition by oxometalate inhibitors of
phosphatases has involved kinetic studies19,121,174,271-280
as well as structural studies.112-127
(8)
Crans et al.
(OH)2}n solubility product unless the total concentration of V(IV) drops below 10-6 M. The predicted
solubility of V(IV) based solely on the positively
charged species ([VO(H2O)5]2+, [VO(H2O)4(OH)]+, and
[{VO(OH)}2]2+) is significantly less than the observed
soluble V(IV). The missing components of V(IV) are
a negatively charged monomeric and dimeric anionic
species ([VO(OH)3]- and [(VO)2(OH)5]-). Concentrations of positively charged species under these conditions are extremely low (for example, with a saturated V(IV) solution at pH 7.4 the expected nanomolar
concentrations of V(IV) species are calculated to be
[VO(H2O)5]2+ ) 1, [VO(H2O)4(OH)]+ ) 30, [{VO(OH)}2]2+ ) 0.8, [(VO)2(OH)5]- ) 350 000, and
[VO(OH)3]- ) 20 000), suggesting that indeed negatively charged species play a key role in biological
systems under anaerobic noncomplexing conditions.
The formula [VO(OH)3]- implies a four-coordinate
species, and the spectroscopic signature of this species was investigated using UV-visible and EPR
spectroscopy to confirm the geometry.289 The results
were not found to be consistent with four- or fivecoordinate species but, rather, with a six-coordinate
species. Neither potentiometry nor other spectroscopic methods employed to date can provide information on the exact number of water molecules
associated with the metal ion in solution. Presumably, two water molecules are associated with the
metal ion making the stoichiometry of this negatively
charged ion [VO(OH)3(H2O)2]-. Note that such a
stoichiometry would readily convert this anion to the
well-known vanadyl cation (47) by three protonation
steps with no changes in the coordination sphere.
The existence of higher oligomeric species (i.e.,
[(VO)2(OH)6]2-, [(VO)4(OH)10]2-, and polymeric {(VO)(OH)3}nn-) was suggested in neutral and basic
Crans et al.
Figure 7. The active site residues in the glutathione-Stransferase fusion protein of domain 1 of pp1B obtained
by NMR studies and modeling studies with a N,N-dimethylhydroxylamine derivative placed in the active site. Redrawn from ref 316. Copyright 1998 The Society of Biological Inorganic Chemistry.
Crans et al.
4.2. Ribonuclease
Ribonucleases are enzymes that catalyze the cleavage of RNA as shown in a simplified manner in
Scheme 6.368 If the cleavage reaction occurs in the
middle of an RNA chain, the ribonuclease is classified
as an endonuclease, and if the cleavage occurs at the
end of the RNA chain it is classified as an exonuclease. Depending on the type of ribonuclease (A or
T1 for example), the RNA cleavage will be specific for
Crans et al.
mutase present in mammals, the 2,3-phosphoglycerate-dependent and 2,3-phosphoglycerate-independent forms, show differential affinities for vanadate
with the cofactor dependent form responding most
potently to vanadate.395 Cofactor-independent forms
of phosphoglycerate mutases are not inhibited by
vanadate at the 0.1 mM level, whereas the reversible
competitive inhibition yields an apparent Ki value of
15 nM for the cofactor-dependent phosphoglycerate
mutase from E. coli. The structure of E. coli cofactordependent phosphoglycerate mutase, complexed with
vanadate, has been determined to a resolution of 1.30
.396 Vanadate is bound in the active site, principally
as a vanadate trimer and not as monomeric vanadate
as reported previously.120,397 These phosphomutases
systems, like those of ribonucleases, provide ample
evidence for the applications of vanadate derivatives
in enzymology. Some mechanistic investigations in
these systems are at a level of detail analyzing the
minor differences between a vanadate-protein complex and that of the true transition state complex of
phosphoglucomutase.
4.4. ATPases
ATPases hydrolyze phosphate-anhydride bonds
and have many important roles in biology in cellular
energy metabolism. Since there are many different types of ATPases, a wide range of affinities for vanadate can be anticipated and are
observed.2,6,12,60,398,399 The effect of vanadate on some
ATPases is very dramatic with a nanomolar inhibition constant having been reported for the Na+, K+
ATPases, compared to the millimolar inhibition
constants for F1-ATPases.398 The ATPase enzymes
include many membrane enzymes, and although
many details are known with regard to the biochemical mechanisms of some ATPases, inhibitory studies
assume that vanadate acts as a phosphate analogue
and presumably inhibits the enzymes as a transition
state analogue for the phosphoryl group transfer.400-402
Information on the formation of stable protein-ViMgADP complexes similar to the protein-Pi-ADP
transition state has been described in detail elsewhere.60 Often transport inhibition in the presence
of vanadate is interpreted as an effect on an ATPase
and is used as a test of how a biological system
responds to an ATPase inhibitor.400-402 The reader
is referred to the general literature and reviews
articles for more information on this topic.2,6,12,60,399
Crans et al.
5.1. Amavadine
Amavadine is a metal complex that contains two
equivalents of ligand and one equivalent of vanadium. The H3hidpa ligand binds V(IV) to form amavadine ([V(hidpa)2]2-) (70),7,440 which is currently the
only documented siderophore-like ligand that binds
vanadium with greater affinity than any other metal
ion (log K2 ) 23).57 The coordinating functionalities
Crans et al.
types of amavadine crystals isolated from Am. muscaria have been grown in the presence of phosphoric
acid and Ca2+ ions, respectively (Figure 11).467 These
structures were solved and not only confirmed the
proposed structure but also documented the chirality
of the amavadine structure both in the neutral and
in the anionic form. Amavadine possesses five chiral
centers: four chiral carbons all have S configuration
and the fifth chiral center is generated by the manner
in which the ligands wrap around the vanadium.
Spectroscopic studies have shown that as isolated,
the natural product contains equal amounts of the
and the forms of amavadine.451 The crystal
structure of the vanadium(IV) complex of meso-2,2(hydroxyimino)dibutyric acid (76) was found to have
similar structural features.58
Figure 11. ORTEP view of a portion of the lattice of [Ca(H2O)5][-V((S,S)-hidpa)2]2H2O. Reproduced with permission from ref 467. Copyright 1999 Wiley-VCH.
CH4 9
8 CH3CO2H
K S O ,HOTf
2 2
(9)
The Garner group is seeking an increased understanding of the amavadine system by investigating
other related metal ion complexes with disubstituted
hydroxylimino ligands. As a result, a number of
different metal complexes have been reported, of
which the titanium,441,470 molybdenum,442,443,471 niobium,444 and tantalum444 systems will be briefly
described here. The titanium system was found to
have a structure similar to amavadine; however,
allowing a solution of the complex to stand for 2
weeks resulted in changes in the NMR spectrum; this
result indicates that solvent molecules do interact
with the hydrophilic equatorial surface of the com-
5.2. Siderophores
Siderophores provide a number of different coordination environments and a wide range of formation
constants with Fe(II) and Fe(III) as well as the
accompanying changes in complex lability.439 The fact
that Fe(III) is bound more tightly and that the Fe(II)
siderophore complex is more labile can explain how
the Fe(III) form can be taken up, the Fe(III)-complex
recognized and transported inside the cell where a
redox switch is used to release the Fe(II) inside the
cell.439 Despite the obvious biological implications if
vanadium is taken up in the place of iron by siderophores, most of the studies have focused on the
characterization of new types of V-complexes with
siderophores and little quantitative information is
available. At this time, the biological role may be
explored by examining the effects of V-compounds on
siderophore mediated-iron transport and has therefore been described in this review.
Crans et al.
features in all of these 2:2 complexes is the diamondcore [V-O-V-O] motif predominant in V(V) alkoxide
chemistry. The V(V)-citrate complexes tend to be fivecoordinate, with the exception of the six-coordinate
peroxo citrate complexes, while the V(IV)-complexes
are generally six-coordinate. Changes in pH did not
significantly affect the structure of the complex ion,
aside from the protonation state(s) of the terminal
carboxylates, in either the V(V)480- or V(IV)485-citrate
complexes. However, some of these complexes were
found to convert to each other in solution when the
pH was changed.480
Aqueous speciation studies of the vanadate(V)citrate found primarily 2:1 complexes.489 In light of
the structural results and the closeness of the fit, one
species, assumed to be a 1:1 complex, may be a 2:2
species.489 The vanadyl(IV)-citrate speciation system
indicated the presence of primarily 2:2 complexes,
although below pH 4.5 1:1 species were detected
while 1:2 (metal:ligand) species could be detected
above pH 8.
A close relative of citrate, homocitrate, is a component of the cofactor in the nitrogenases (Nases),
including V-nitrogenase (vide infra). Only one structure has been crystallographically determined for a
V(V)-complex with homocitrate (91).490 Discussion of
Nase and model compounds is presented below in the
section on low-valent vanadium chemistry.
complexes, 15N NMR spectra suggest the hydroxamate to be monoanionic and 13C NMR spectra indicate
that the carbonyl oxygen is weakly coordinated. The
log Kf values obtained (ca. 1-4)496 were much lower
than those obtained in more acidic solutions.497,498
Speciation studies on alkylhydroxamic acid-vanadate complexes give log Kf values on the order of
7-38 depending on the pH.499,500 Cinnamoylhydroxamate was shown to form 1:1 and 1:2 complexes with
V(V) at 1.8 M HCl in 2-methyl-4-pentanone; the
stepwise formation constants were given and an
octahedral coordination geometry was suggested.501
The first structurally characterized V-hydroxamate
complexes were the V(V)-complexes of benzohydroxamic acid (92),502 the dihydroxamate N,N-dihydroxyN,N-diisopropylheptanediamide,502 and N-phenylbenzohydroxamic acid (93).503 All the complexes are
six-coordinate and the ligands are monoanionic; the
benzohydroxamate and N-phenylbenzohydroxamate
complexes are monomeric, while the dihydroxamate
ligand yielded a 2:2 complex. A series of salicylhydroxamate V(V)-complexes with tridentate ternary
ligands were prepared; the derivative with N(salicylideneaminato)-N-(2-hydroxyethyl)ethylenediamine was crystallized (94).246 This complex and
its congeners contain a dianionic salicylhydroxamate
ligand (i.e., a hydroximate), and a strong hydrogen
bond was observed between the phenolic hydroxyl
group and the deprotonated imino nitrogen.246 A
trinuclear vanadium(V) complex, [VO(shi)(OCH3)]3
(95), was prepared from VO(acac)2 or VCl3, salicylhydroxamic acid, and NaOCH3 in methanol and
structurally characterized. Each trianionic ligand is
tetradentate with two donor atoms bound to each
vanadium center to generate a metallocrown structure.504 Similar metallocrown structures have also
been prepared from modified salicylhydroxamic acids.505 The first mixed hydroxamate/hydrazone V(V)complex contained a monoanionic salicylhydroxamate
ligand.506 Other structurally characterized V(V) ternary complexes include 4-(2-(salicylideneamino)ethyl)-imidazole/salicylhydroximate,250 benzohydroxamic acids/N-salicylideneglycine and N-(2-carboxyphenyl)salicylideneamine,507 and N-phenylbenzohydroxamic acid/acetylacetone benzoylhydrazonato and
salicylidene-L-alaninato.508
To date, no crystallographic data for V(IV) hydroxamate complexes are available, although solution
studies indicate that V(IV) hydroxamate complexes
Crans et al.
6.1. Tunicates
Tunicates (ascidians or sea squirts) are invertebrate marine organisms and, depending on the species, accumulate vanadium in their blood. The V-containing species were first discovered in 1911 by
Henze,4,7,522 and since then, bioinorganic chemists
and biological scientists have been interested in these
animals that can concentrate vanadium from seawater (10-8 M) in various cells of the tunicate (Table
2). Bioinorganic chemists have investigated the form
and storage of vanadium in these organisms, its
redox conversion from V(V) in seawater to V(IV) and
V(III), and also how and why vanadium is accumulated. Although much is known about most of
Table 2. Concentration of Vanadium (M) in Ascidian
Tissuesa
species
Ascidia gemmata
A. ahodori
A. sydneiensis
Phallusia
mammillata
Ciona intestinalis
Styela plicata
Halocynthia roretzi
H. aurantium
a
serum
blood
cells
ND
1000
50
ND
347 200
59 900
12 800
19 300
700
600
Stolidobranchia
5
1
1
10
1
4
2
2
2
3
1
ND
3
7
4
tunic mantle
branchial
Phlebobranchia
ND
ND
ND
2400 11 200 12 900
60
700
1.400
30
900
2.900
3
700
Crans et al.
Figure 13. Sulfur K-edge XAS spectra of (a) (s) the highvalent portion of the blood cell spectrum; (- - -), the fit
to the spectrum; and (- - -) the Gaussian components of the
fit to the spectrum. (b) The second derivatives of (s) the
blood cell sulfur K-edge XAS spectrum and (- - -) the fit
to the spectrum. Reprinted from ref 557. Copyright 1995
American Chemical Society.
Crans et al.
2VIVO(salen) + 4H+ f
VVO(salen)+ + VIII(H2salen) + H2O (10)
VVO(salen)+ + VIII(H2salen) + 4Cl- f
VIVO(salen) + VIVCl2(salen) + 2HCl (11)
including a dinuclear V(V) catecholate complex [Et3NH]2[VO2(3,5-dtbc)]2 (110).574,575 Because of their rich
properties, these systems have versatile reactivity
patterns, some of which will be described as model
systems for the nitrogenase reaction.591-594
Representative examples of structurally characterized vanadium-sulfate compounds are shown (111117). Most of the structural studies of vanadium-
Crans et al.
7. Vanadium Nitrogenase
7.1. Nitrogenases
The bacterial enzyme nitrogenase (Nase) is capable
of catalyzing the reduction of atmospheric N2 to NH3
and is responsible for cycling about 108 tons of N per
year from the atmosphere to the soil.27 Nase is a
metalloprotein containing Mo and Fe, V, and Fe or
Fe only as the metal cofactors. The enzymatic N2
fixation occurs at ambient temperature and 0.8 atm
N2 pressure; modeling this process remains one of the
great challenges for bioinorganic chemists. The fascination with this protein can be explained, in part,
because of the several protein and inorganic components, and in part because of the complex chemical
reaction that is catalyzed by this enzyme. The overall
reaction of the common molybdenum-containing Nase
is currently believed to be as shown in eq 12. Some
variation in the stoichiometry of e-, H+, and MgATPs
and the overall products of the reaction exists depending on the specific conditions under which the
reaction occurs and which Nase is used. A greater
amount of H2 is generated for each NH3 produced and
more ATP is consumed during the reduction of N2
by the vanadium nitrogenase (V-Nase) (eq 13) as
compared to the molybdenum nitrogenase (MoNase) (eq 12).27 Whether the exact number of MgATP
required for the V-Nase reaction is 4027,30 or 2428-30
can presumably be traced to the difficulties in the
preparation and purification of the enzyme. The
enzymes carry out their function under anaerobic
conditions and have developed elaborate protection
mechanisms to exclude oxygen from the active site
and the redox active cofactors.
Crans et al.
incorporates two binding sites for MgATP. A schematic illustration of functional aspects of the complex
is shown in Figure 16. Investigations into the enzymology of this process have recently been reexamined
using ATP-analogues after it was recognized that this
protein contains the peptide fold common for nucleotide binding proteins.623,624
The FeV component has a R222 subunit structure,
and contains two P clusters at the R and subunit
interface. The P cluster is believed to transfer electrons to the FeM-cluster in all three Nases. The M
cluster in the V-Nase is commonly referred to as
FeVco and is the proposed binding site for N2. Two
FeVco clusters are located in the R subunits (Figure
16). The FeVco cluster can be described as the
catalytic cofactor since it is responsible for the
conversion of N2 to NH3 (vide infra). The exact
structure of the VFe protein is still not known, but
all studies so far strongly suggest that it is analogous
to the Mo-Nase.625 A second V-Nase has been found
in Az. vinelandii, which lacks the R subunit and half
of the P cluster, but is active nonetheless.626,627 Thus
far, all the clusters have only been assembled by
biosynthesis.
In the absence of a crystal structure of a VFe
protein, the detail of the environment of the Vbinding site has been provided from EXAFS studies
of proteins isolated from Azotobacter chroococcum
(Ac1v)628 and Az. vinelandii (Av1v).629 The V-atom was
in an environment similar to that observed for the
V-atom in the model compound [Me4N][VFe3S4Cl3(dmf)3] (129).630,631 XANES and EXAFS studies of
thionine-oxidized and dithionite-reduced forms of the
Az. chroococcum Nase suggest a vanadium oxidation
state between +II and +IV.628,632 EPR studies in
reduced Az. vinelandii V-Nase give g ) 5.5, which
corresponds to an overall S ) 3/2 spin for the cluster
in the protein.51,633
Temperature has been shown to play a role not
only with regard to the activity of the Nases, but also
in the expression of the enzymes themselves. Upon
increasing the temperature from 30 to 40 C, the
amount of N2H4 produced by the Az. chroococcum
V-Nase increased by a factor of 15, whereas the
amount of NH3 produced remained constant.634 Increasing the temperature to 50 C resulted in a
further 3-fold increase in N2H4 production and a
7-fold decrease in NH3 production.635 These reversible
effects were interpreted as temperature-induced conformational changes, which at high temperatures,
may allow FeVco to reduce N2 to N2H4, but not to
NH3.51,634 In the same system, V-Nase was shown
to have 10 times the activity of the Mo-Nase when
the temperature was lowered to 5 C.636 In the Az.
vinelandii enzyme system, at 30 C, Mo was found
to repress both the vnf and anf operons, less so at 20
C, and not at all at 14 C.637 Furthermore, vanadium
was shown to repress the anf operon at 30 C, but
not at 20 C; this suggests that Az. vinelandii can
produce all three Nases between 14 and 20 C,
depending on the availability of Mo and V. These
results suggest that Nases found in cooler climates
may favor the presence of V-Nases.
Figure 17. Structure of the FeMo-cofactor of dithionitereduced A. vinelandii MoFe-protein. Redrawn from ref 641.
Copyright 2002 American Association for the Advancement
of Science.
These clusters also catalyze the reduction of phenylhydrazine to NH3 and aniline. Attempts to characterize intermediates by the isolation and characterization of complexes with hydrazine have led to
the report of the hydrazine vanadium single cubane
complex (Me4N)[(PhHNNH2)(bpy)VFe3S4Cl4] (140).655
A related cluster (141) reacts to form (142) and
subsequently (143).648,649 The S-bridged structure of
[(Bpz)2V2Fe6S9(SH)2]4- (143) has crystallographically
imposed C2 symmetry. The core exhibits a bridging
pattern [V2Fe6(2-S)2(3-S)6(6-S)], which reduces to
two cuboidal VFe3(3-S)3 fragments that share a
common bridging atom (the 6-S atom) and are
externally bridged by two 2-S atoms. Cyclic voltammetry experiments on the [VFe3S4]2+ cluster,
[(HBpz)3VFe3S4(LS3)]2- (136), demonstrated the stability of this system by reversible conversion among
the 1+, 2+, and 3+ oxidation states of the clusters.
Double-cubane clusters such as [V2Fe6S8Cl4(C2H4S2)2]4- (135)652 have been reported and their
reactivity investigated. When this cluster reacts with
HSCH2CH2SH in dmso, a single cubane-cluster was
formed (133). Single and double cubane clusters in
multiple oxidation states were isolated from [(HBpz)3-
Crans et al.
7.5. Homocitrate
Homocitrate is a noncovalent cofactor and a structural component of FeMoco. It is presumably also a
component of FeVco and the iron-iron (FeFeco)
clusters, because nifV is required for full functionality
of all three Nases.663 A 49V labeled study showed that
homocitrate is required for the V-precursor transfer
of FeVco from VnfX to nif-apodinitrogenase.664 The
dinuclear V(V) species [K2(H2O)5][(VO2)2(R,Shomocitrate)2]H2O represents the first synthetic and
structural characterization of a transition metal
homocitrate complex and may in fact be an early
precursor in the biosynthesis of FeVco.490
For comparison, several V(V)-citrate complexes
have been reported.480,483,490 The dinuclear V(V)citrate complex, K2[V(O)2(C6H6O7)]24H2O, demonstrated the coordination mode found in the Nase
cofactor.483 See the siderophore section of this review
for additional details on V-citrate complexes.
Crans et al.
(15)
(16)
(14)
Substituted hydrazines form dinuclear V(III)complexes in which the hydrazine bridges the two
V-atoms (164).680 In mononuclear complexes 167171, the two hydrazine groups are generally coordinated head-on in a trans configuration.680 While most
complexes have the hydrazine group coordinating
head-on (such as 171681) some are thought to coordinate side-on, 168 and 169.682 Although reactivity
studies need to be examined for many of these
of simple vanadium salts and peroxovanadium compounds on protein tyrosine phosphatases and insulin
receptor phosphorylation have been useful agents in
studies probing the phosphorylated insulin receptor.319 These studies were important because they
demonstrated that hydrolysis of the peroxovanadium
compound did not explain the observed effects observed in cells or in animal model systems. A large
class of compounds based on V(IV) chelate complexes726,740 have been extensively studied mainly by
the Sakurai and Orvig/McNeill groups, respectively.64,729,740,741 The effects of bis(picolinato)oxovanadium(IV)740,742 and the effects of bis(maltolato)oxovanadium(IV),727 have been described in great
detail and have also been reviewed.64 In addition,
recent reports have documented the effects of certain
V(V)-complexes in animals.66,67,743 One class of V(V)complexes are the dipicolinate vanadate complexes,
and the chemistry of these V-complexes have been
examined in detail.66,67 Another class of V(V)-complexes are formed with hydroxamate ligands, and
these compounds have been prepared in situ.743 These
studies counter earlier reports of inactive V(V)complexes derived from maltol.728 A few insulin
enhancing V(III)-complexes have been reported68
suggesting that the vanadium oxidation state does
not seem to be as critical as previously believed.
Reviews describing the effects of a number of Vcompounds tested in animals show how the number
of compounds is rapidly increasing.64,737 The insulin
mimetic effect of V-complexes has now been reported
to extend from V to other metal complexes including
Cr,744 Zn,745-747 and Co complexes.748
Because the insulin signaling system is exceptionally complex, we have yet to understand all of its
intricate details. As such, it is not reasonable to
expect the action of effective insulin-enhancing agents
to be readily understood. Indeed, to date, no single
target can explain all the observed insulin-like effects
of V-compounds.323,327,749-751 The insulin-like effect of
V-compounds likely involves a protein tyrosine phosphatase.13,175,751 Specifically, vanadate is generally
believed to exert its insulin enhancing effect through
competitive inhibition of regulatory protein phosphatases, with a major candidate being phosphatase
1B (pp1B). This phosphatase is the first phosphatase
in the insulin regulatory cascade and is particularly
sensitive to inhibition by V-compounds. Inhibition of
pp1B in a knock-out mouse led to increased insulin
sensitivity and obesity resistance752 and with antisense RNA led to increases in insulin-dependent
signaling in the ob/ob mouse.753 However, the potency
of the competitive inhibition of phosphatases by
V-compounds depends on the structure of the complex, the oxidation state of the metal ion, and the
nature of the phosphatase13,272,314,316,319,754,755 and differs by several orders of magnitude, allowing for
selectivity of V-compounds in vivo. Some V-compounds such as peroxovanadate irreversibly oxidize
the catalytic cysteine in protein phosphatases,319
while simple vanadate salts319,739 (and presumably
[VO(malto)2]) do not undergo similar redox chemistry,
which may explain their smaller effect on the autophosphorylation of the IR.175-177 It is likely that other
Crans et al.
attainable in vivo concentrations of V(IV). The coordination complexes that form between V(IV) and
GSH have been described by several groups,300,301,761-764
and the specific species that form are somewhat
controversial depending on pH, and the metal-toligand ratio.300,301,761,763,764 Surprisingly, little structural information is available on metal ion-GSH
complexes,765 although one related V(IV)-complex
with cysteine methyl ester766 has been reported. A
variety of spectroscopic techniques have been used
to examine V-complexes. Depending on the conditions, eight to nine different V(IV) species have been
observed by EPR spectroscopy in the presence of 100fold excess of GSH.300,301 Because GSH forms even
weaker complexes with V(V) than with V(IV) with
GSH, and because GSH will undergo oxidation while
V(V) is reduced, this reaction is much less understood.
The role of oxidative stress in the etiology of
diabetes has recently been proposed to be involved
in both the origin of the disease and increasing
secondary complications.767 The role of oxidative
stress caused by glucose toxicity and the resulting
production of free radicals, especially in the pancreas,
has been proposed to be a major cause of the
development of insulin resistance in both type 1768
and type 2 diabetics.769-771 This is especially important in patients with diabetic complications. Since
antioxidant therapy is currently being investigated
to decrease the loss of insulin sensitivity768 it is likely
that some of the effects of metal complexes, whether
they be beneficial or toxic effects, may be due to redox
regulation. In rats with STZ-induced diabetes, vanadate administration was shown to decrease GSH
while the levels of GSH reductase and oxidized GSH
remained unchanged.772 In comparison with insulin,
vanadate was only partially able to control the
impaired antioxidation system of diabetic rats. In the
same model system, both vanadate and insulin
improved lipid metabolism as monitored by total lipid
levels, triglyceride, and lipogenic enzymes.773 In
examining different tissues, vanadate administration
was also found to have antioxidant properties when
the activities of antioxidant enzymes catalase, superoxide dismutase, and GSH peroxidase were measured.774 The above results suggest that antioxidant
effects of the oral administration of vanadate have
been established in diabetic rats and support the
hypothesis that changes in cellular GSH metabolism
are connected to the insulin-enhancing properties of
transition metal complexes.
Current recognition of both the negative and the
positive effects of oxidation-reduction reactions in
biology has led to the cellular redox state being
increasingly investigated. Various examples of protein and metabolism regulation have been described,775-778 including the proposal that cellular
oxidation mediates insulin resistance.767 Decreased
glutathione (GSH) levels have been implicated in the
lowered resistance of diabetics to oxidative stress.779
Complications of diabetes have been correlated with
reduced GSH levels, suggesting that antioxidants
such as GSH can protect against diabetic complications.780 The vanadate-stimulated NAD(P)H oxida-
9. Summary
V-compounds and vanadium-containing proteins
have shown a wide range of properties and reactivities, some small-molecule based while others provide
the needed structure or function in a biomolecule.
Various aspects of this topic have been described in
more than 100 reviews over the past decade. It was
therefore beyond the scope of this review to be
comprehensive. Indeed, the readers are referred
elsewhere for more information in the various areas,
and key reviews are listed to provide the novice
Crans et al.
10. Acknowledgments
DCC thanks the Institute for General Medicine at
the National Institutes of Health for funding.
12. References
(1) Nielsen, F. H.; Uthus, E. O. In Vanadium in Biological Systems;
Chasteen, N. D., Ed.; Kluwer Academic Publishers: Boston,
1990.
(2) Kustin, K.; McLeod, G. C.; Gilbert, T. R.; Briggs, L. B. R. T.
Struct. Bonding 1983, 53, 139.
(3) Kustin, K.; Robinson, W. E.; Smith, M. J. Invertebr. Reprod. Dev.
1990, 17, 129.
(4) Michibata, H.; Sakurai, H. In Vanadium in Biological Systems;
Chasteen, N. D., Ed.; Kluwer Academic Publishers: Boston,
1990.
(5) Smith, M. J.; Ryan, D. E.; Nakanishi, K.; Frank, P.; Hodgson,
K. O. Met. Ions Biol. Syst. 1995, 31, 423.
(6) Wever, R.; Kustin, K. Adv. Inorg. Chem. 1990, 35, 81.
(7) Slebodnick, C.; Hamstra, B. J.; Pecoraro, V. L. Struct. Bonding
1997, 89, 51.
(8) Taylor, S. W.; Kammerer, B.; Bayer, E. Chem. Rev. 1997, 97,
333.
(9) Rehder, D. Coord. Chem. Rev. 1999, 182, 297.
(10) Michibata, H.; Yamaguchi, N.; Uyama, T.; Ueki, T. Coord. Chem.
Rev. 2003, 237, 41.
(11) Michibata, H.; Uyama, T.; Ueki, T.; Kanamori, K. Microsc. Res.
Technol. 2002, 56, 421.
(12) Chasteen, N. D. Struct. Bond 1983, 53, 105.
(13) Gresser, M. J.; Tracey, A. S.; Stankiewicz, P. J. Adv. Prot.
Phosphatases 1987, 4, 35.
(14) Gresser, M. J.; Tracey, A. S. In Vanadium in Biological Systems;
Chasteen, N. D., Ed.; Kluwer Academic Publishers: Boston,
1990.
(15) Vanadium in Biological Systems; Chasteen, N. D., Ed.; Kluwer
Academic Publishers: Dordrecht, The Netherlands, 1990.
(16) Rehder, D. Angew. Chem., Int. Ed. Engl. 1991, 30, 148.
(17) Crans, D. C. Comments Inorg. Chem. 1994, 16, 1.
(18) Rehder, D. Met. Ions Biol. Syst. 1995, 31, 1.
(19) Crans, D. C.; Keramidas, A. D.; Drouza, C. Phosphorus, Sulfur,
Silicon Relat. Elem. 1996, 109-110, 245.
(20) Tracey, A. S.; Crans, D. C. ACS Symp. Ser. 1998, 711.
(21) Van Pee, K.-H.; Keller, S.; Wage, T.; Wynands, I.; Schnerr, H.;
Zehner, S. Biol. Chem. 2000, 381, 1.
(22) Butler, A. In Bioinorganic Catalysis, 2nd ed.; Reedijk, J.,
Blouwman, E., Eds.; Marcel Dekker: New York, 1999.
(23) Butler, A. Curr. Opin. Chem. Biol. 1998, 2, 279.
(24) Wever, R.; Barnett, P.; Hemrika, W. Transition Met. Microb.
Metab. 1997, 415.
(25) Vilter, H. Met. Ions Biol. Syst. 1995, 31, 325.
(26) Butler, A.; Walker, J. V. Chem. Rev. 1993, 93, 1937.
(27) Eady, R. R. Coord. Chem. Rev. 2003, 237, 23.
(28) Rehder, D. J. Inorg. Biochem. 2000, 80, 133.
Crans et al.
(78) Nour-Eldeen, A. F.; Craig, M. M.; Gresser, M. J. J. Biol. Chem.
1985, 260, 6836.
(79) Drueckhammer, D. G.; Durrwachter, J. R.; Pederson, R. L.;
Crans, D. C.; Daniels, L.; Wong, C.-H. J. Org. Chem. 1989, 54,
70.
(80) Crans, D. C.; Simone, C. M.; Blanchard, J. S. J. Am. Chem. Soc.
1992, 114, 4926.
(81) Crans, D. C.; Marshman, R. W.; Nielsen, R.; Felty, I. J. Org.
Chem. 1993, 58, 2244.
(82) Prandtl, W.; Hess, L. Z. Anorg. Chem. 1913, 82, 103.
(83) Orlov, N. F.; Voronkov, M. G. Bull. Acad. Sci., USSR Chem. Sci.
1959, 899.
(84) Mittal, R. K.; Mehrotra, R. C. Z. Anorg. Allg. Chem. 1964, 327,
311.
(85) Lachowicz, A.; Hobold, W.; Thiele, K.-H. Z. Anorg. Allg. Chem.
1975, 418, 65.
(86) Priebsch, W.; Rehder, D. Inorg. Chem. 1990, 29, 3013.
(87) Crans, D. C.; Felty, R. A.; Miller, M. M. J. Am. Chem. Soc. 1991,
113, 265.
(88) Hillerns, F.; Olbrich, F.; Behrens, U.; Rehder, D. Angew. Chem.,
Int. Ed. Engl. 1992, 31, 447.
(89) Crans, D. C.; Chen, H.; Felty, R. A. J. Am. Chem. Soc. 1992,
114, 4543.
(90) Caughlan, C. N.; Smith, H. M.; Watenpaugh, K. Inorg. Chem.
1966, 5, 2131.
(91) Feher, F. J.; Walzer, J. F. Inorg. Chem. 1991, 30, 1689.
(92) Henderson, R. A.; Hughes, D. L.; Janas, Z.; Richards, R. L.;
Sobota, P.; Szafert, S. J. Organomet. Chem. 1998, 554, 195.
(93) Crans, D. C.; Felty, R. A.; Anderson, O. P.; Miller, M. M. Inorg.
Chem. 1993, 32, 247.
(94) Toscano, P. J.; Schermerhorn, E. J.; Dettelbacher, C.; Macherone,
D.; Zubieta, J. J. Chem. Soc., Chem. Commun. 1991, 933.
(95) Crans, D. C.; Felty, R. A.; Chen, H.; Eckert, H.; Das, N. Inorg.
Chem. 1994, 33, 2427.
(96) Kempf, J. Y.; Maigret, B.; Crans, D. C. Inorg. Chem. 1996, 35,
6485.
(97) Mondal, S.; Dutta, S.; Chakravorty, A. J. Chem. Soc., Dalton
Trans. 1995, 1115.
(98) Rath, S. P.; Rajak, K. K.; Mondal, S.; Chakravorty, A. J. Chem.
Soc., Dalton Trans. 1998, 2097.
(99) Rath, S. P.; Rajak, K. K.; Chakravorty, A. Inorg. Chem. 1999,
38, 4376.
(100) Gresser, M. J.; Tracey, A. S. J. Am. Chem. Soc. 1985, 107, 4215.
(101) Tracey, A. S.; Gresser, M. J. Proc. Natl. Acad. Sci., U.S.A. 1986,
83, 609.
(102) Elvingson, K.; Crans, D. C.; Pettersson, L. J. Am. Chem. Soc.
1997, 119, 7005.
(103) Heath, E.; Howarth, O. W. J. Chem. Soc., Dalton Trans. 1981,
1105.
(104) Day, V. W.; Klemperer, W. G.; Yaghi, O. M. J. Am. Chem. Soc.
1989, 111, 4518.
(105) Zavalu, P. Y.; Zhang, F.; Whittingham, M. S. Acta Crystallogr.
1997, C53, 1738.
(106) Tracey, A. S.; Gresser, M. J.; Liu, S. J. Am. Chem. Soc. 1988,
110, 5869.
(107) Geraldes, C. F. G. C.; Castro, M. M. C. A. J. Inorg. Biochem.
1989, 37, 213.
(108) Boyd, D. W.; Kustin, K.; Niwa, M. Biochim. Biophys. Acta 1985,
827, 472.
(109) Sakurai, H.; Goda, T.; Shimomura, S.; Yoshimura, T. Biochem.
Biophys. Res. Commun. 1982, 104, 1421.
(110) Williams, P. A. M.; Etcheverry, S. B.; Baran, E. J. J. Inorg.
Biochem. 1996, 61, 285.
(111) Williams, P. A. M.; Baran, E. J. J. Inorg. Biochem. 1992, 48, 15.
(112) Ladner, J. E.; Wladkowski, B. D.; Svensson, L. A.; Sjolin, L.;
Gilliland, G. L. Acta Crystallogr. 1997, D53, 290.
(113) Rupert, P. B.; Massey, A. P.; Sigurdsson, S. T.; Ferre-DAmare,
A. R. Science 2002, 298, 1421.
(114) Davies, D. R.; Interthal, H.; Champoux, J. J.; Hol, W. G. J. J.
Mol. Biol. 2002, 324, 917.
(115) Smith, C. A.; Rayment, I. Biochemistry 1996, 35, 5404.
(116) Holtz, K. M.; Stec, B.; Kantrowitz, E. R. J. Biol. Chem. 1999,
274, 8351.
(117) Bond, C. S.; White, M. F.; Hunter, W. N. J. Mol. Biol. 2002, 316,
1071.
(118) Kollmar, M.; Durrwang, U.; Kliche, W.; Manstein, D.; Kull, F.
J. EMBO J. 2002, 21, 2517.
(119) Stuckey, J. A.; Schubert, H. L.; Fauman, E. B.; Zhang, Z.-Y.;
Dixon, J. E.; Saper, M. A. Nature 1994, 370, 571.
(120) Lindqvist, Y.; Schneider, G.; Vihko, P. Eur. J. Biochem. 1994,
33, 139.
(121) Zhang, M.; Zhou, M.; Van Etten, R. L.; Stauffacher, C. V.
Biochemistry 1997, 36, 15.
(122) Fauman, E. B.; Yuvaniyama, C.; Schubert, H. L.; Stuckey, J.
A.; Saper, M. A. J. Biol. Chem. 1996, 271, 18780.
(123) Barford, D.; Flint, A. J.; Tonks, N. K. Science 1994, 263, 1397.
(124) Schneider, G.; Lindqvist, Y.; Vihko, P. EMBO J. 1993, 12, 2609.
(125) Su, X.-D.; Taddel, N.; Stefani, M.; Ramponi, G.; Nordlund, P.
Nature 1994, 3670, 575.
Crans et al.
(277) Posner, B. I.; Faure, R.; Burgess, J. W.; Bevan, A. P.; Lachance,
D.; Zhang-Sun, G.; Fantus, I. G.; Ng, J. B.; Hall, D. A.; Soo Lum,
B.; Shaver, A. J. Biol. Chem. 1994, 269, 4596.
(278) Reiter, N. J.; White, D. J.; Rusnak, F. Biochemistry 2002, 41,
1051.
(279) Foster, J. D.; Young, S. E.; Brandt, T. D.; Nordlie, R. C. Arch.
Biochem. Biophys. 1998, 354, 125.
(280) Ohlson, J. T.; Wilson, I. B. Biochim. Biophys. Acta 1974, 350,
48.
(281) Hengge, A. C.; Zhao, Y.; Wu, L.; Zhang, Z.-Y. Biochemistry 1997,
36, 7928.
(282) Zhang, Z.-Y.; Wang, Y.; Dixon, J. E. Proc. Natl. Acad. Sci. U.S.A.
1994, 91, 1624.
(283) Deng, H.; Callender, R.; Huang, Z.; Zhang, Z.-Y. Biochemistry
2002, 41, 5865.
(284) Lopez, V.; Stevens, T.; Lindquist, R. N. Arch. Biochem. Biophys.
1976, 175, 31.
(285) Francavilla, J.; Chasteen, N. D. Inorg. Chem. 1975, 14, 2860.
(286) Henry, R. P.; Mitchell, P. C.; Prue, J. E. J. Chem. Soc., Dalton
Trans. 1973, 1156.
(287) Rossotti, F. J. C.; Rossotti, H. S. Acta Chem. Scand. 1955, 9,
1177.
(288) Komura, A.; Hayashi, M.; Imanaga, H. Bull. Chem. Soc. Jpn.
1977, 50, 2927.
(289) Iannuzzi, M. M.; Rieger, P. H. Inorg. Chem. 1975, 14, 2895.
(290) Britton, H. T. S.; Welford, G. J. Chem. Soc. 1940, 758.
(291) Lingane, J. J.; Meites, L., Jr. J. Am. Chem. Soc. 1947, 69, 1021.
(292) Ostrowetsky, S. Bull. Soc. Chim. Fr. 1964, 1012.
(293) Pessoa, J. C.; Boas, L. F. V.; Gillard, R. D.; Lancashire, R. J.
Polyhedron 1988, 7, 1245.
(294) Pessoa, J. C.; Boas, L. F. V.; Gillard, R. D. Polyhedron 1990, 9,
2101.
(295) Chasteen, N. D.; Grady, J. K.; Holloway, C. E. Inorg. Chem. 1986,
25, 2754.
(296) Alberico, E.; Micera, G. Inorg. Chim. Acta 1994, 215, 225.
(297) Kiss, T.; Kiss, E.; Micera, G.; Sanna, D. Inorg. Chim. Acta 1998,
283, 202.
(298) Buglyo, P.; Kiss, E.; Fabian, I.; Kiss, T.; Sanna, D.; Garriba, E.;
Micera, G. Inorg. Chim. Acta 2000, 306, 174.
(299) Micera, G.; Sanna, D.; Dessi, A.; Kiss, T.; Buglyo, P. Gazz. Chim.
Ital. 1993, 123, 573.
(300) Pessoa, J. C.; Tomaz, I.; Kiss, T.; Buglyo, P. J. Inorg. Biochem.
2001, 84, 259.
(301) Pessoa, J. C.; Tomaz, I.; Kiss, T.; Kiss, E.; Buglyo, P. J. Biol.
Inorg. Chem. 2002, 7, 225.
(302) Micera, G.; Sanna, D.; Kiss, E.; Garribba, E.; Kiss, T. J. Inorg.
Biochem. 1999, 75, 303.
(303) Branca, M.; Micera, G.; Dessi, A.; Sanna, D. J. Inorg. Biochem.
1992, 45, 169.
(304) Garribba, E.; Lodyga-Chruscinska, E.; Sanna, D.; Micera, G.
Inorg. Chim. Acta 2001, 322, 87.
(305) Kiss, T.; Buglyo, P.; Sanna, D.; Micera, G.; Decock, P.; Dewaele,
D. Inorg. Chim. Acta 1995, 239, 145.
(306) Buglyo, P.; Kiss, T.; Kiss, E.; Sanna, D.; Garribba, E.; Micera,
G. J. Chem. Soc., Dalton Trans. 2002, 2275.
(307) Ray, W. J., Jr.; Crans, D. C.; Zheng, J.; Burgner, J. W., II; Deng,
H.; Mahroof-Tahir, M. J. Am. Chem. Soc. 1995, 117, 6015.
(308) Crans, D. C.; Bunch, R. L.; Theisen, L. A. J. Am. Chem. Soc.
1989, 111, 7597.
(309) Crans, D. C.; Gottlieb, M. S.; Tawara, J.; Bunch, R. L.; Theisen,
L. A. Anal. Biochem. 1990, 188, 53.
(310) Williams, P. A. M.; Barrio, D. A.; Etcheverry, S. B. J. Inorg.
Biochem. 1999, 75, 99.
(311) Etcheverry, S. B.; Barrio, D. A.; Williams, P. A. M.; Baran, E. J.
Biol. Trace Elem. Res. 2001, 84, 227.
(312) Salice, V. C.; Cortizo, A. M.; Dumm, C. L. G.; Etcheverry, S. B.
Mol. Cell. Biochem. 1999, 198, 119.
(313) Tracey, A. S. J. Inorg. Biochem. 2000, 80, 11.
(314) Krejsa, C. M.; Nadler, S. G.; Esselstyn, J. M.; Kavanagh, T. J.;
Ledbetter, J. A.; Schieven, G. L. J. Biol. Chem. 1997, 272, 11541.
(315) Nxumalo, F.; Tracey, A. S. J. Biol. Inorg. Chem. 1998, 3, 527.
(316) Nxumalo, F.; Glover, N. R.; Tracey, A. S. J. Biol. Inorg. Chem.
1998, 3, 534.
(317) Bhattacharyya, S.; Tracey, A. S. J. Inorg. Biochem. 2001, 85, 9.
(318) Krejsa, C. M.; Schieven, G. L. Environ. Health Perspect. 1998,
106, 1179.
(319) Huyer, G.; Liu, S.; Kelly, J.; Moffat, J.; Payette, P.; Kennedy,
B.; Tsaprailis, G.; Gresser, M. J.; Ramachandran, C. J. Biol.
Chem. 1997, 272, 843.
(320) Boissonneault, M.; Chapdelaine, A.; Chevalier, S. Mol. Cell.
Biochem. 1995, 153, 139.
(321) Trowbridge, I. S.; Thomas, M. L. Annu. Rev. Immunol. 1994,
12, 85.
(322) Faure, R.; Baquiran, G.; Bergeron, J. J. M.; Posner, B. I. J. Biol.
Chem. 1992, 267, 11215.
(323) Shechter, Y.; Goldwaser, I.; Mironchik, M.; Fridkin, M.; Gefel,
D. Coord. Chem. Rev. 2003, 237, 3.
Crans et al.
(462) Ludwig, E.; Hefele, H.; Uhlemann, E.; Weller, F.; Klaeui, W. Z.
Anorg. Allg. Chem. 1995, 621, 23.
(463) Mickler, W.; Moenner, A.; Hefele, H.; Ludwig, E.; Uhlemann,
E. Electrochim. Acta 1996, 42, 421.
(464) Neves, A.; Ceccato, A. S.; Vencato, I.; Mascarenhas, Y. P.;
Erasmus-Buhr, C. J. Chem. Soc., Chem. Commun. 1992, 652.
(465) Ramesh, K.; Lal, T. K.; Mukherjee, R. N. Polyhedron 1992, 11,
3083.
(466) Vergopoulos, V.; Jantzen, S.; Julien, N.; Rose, E.; Rehder, D. Z.
Naturforsch. 1994, B49, 1127.
(467) Berry, R. E.; Armstrong, E. M.; Beddoes, R. L.; Collison, D.;
Ertok, S. N.; Helliwell, M.; Garner, C. D. Angew. Chem., Int.
Ed. Engl. 1999, 38, 795.
(468) Matoso, C. M. M.; Pombeiro, A. J. L.; Da Silva, J. J. R. F.; Da
Silva, M. F. C. G.; Da Silva, J. A. L.; Baptista-Ferreira, J. L.;
Pinho-Almeida, F. ACS Symp. Ser. 1998, 711, 241.
(469) Reis, P. M.; Armando, J.; Silva, L.; da Silva, J. J. R. F.; Pombeiro,
A. J. L. Chem. Commun. 2000, 1845.
(470) Harben, S. M.; Smith, P. D.; Beddoes, R. L.; Collison, D.; Garner,
C. D. Angew. Chem., Int. Ed. Engl. 1997, 36, 1897.
(471) Harben, S. M.; Smith, P. D.; Beddoes, R. L.; Collison, D.; Garner,
D. C. J. Chem. Soc. 1997, 2777.
(472) Baysse, C.; De Vos, D.; Naudet, Y.; Vandermonde, A.; Ochsner,
U.; Meyer, J.-M.; Budzikiewicz, H.; Schafer, M.; Fuchs, R.;
Cornelis, P. Microbiology 2000, 146, 2425.
(473) Allnutt, F. C. T.; Bonner, J. W. D. Plant Physiol. 1987, 85, 746.
(474) Cornish, A. S.; Page, W. J. Appl. Environ. Microbiol. 2000, 66,
1580.
(475) Karpishin, T. B.; Dewey, T. M.; Raymond, K. N. J. Am. Chem.
Soc. 1993, 115, 1842.
(476) Batinic, I.; Birus, M.; Pribanic, M. Croat. Chem. Acta 1987, 60,
279.
(477) Batinic-Haberle, I.; Birus, M.; Pribanic, M. Inorg. Chem. 1991,
30, 4882.
(478) Buglyo, P.; Culeddu, N.; Kiss, T.; Micera, G.; Sanna, D. J. Inorg.
Biochem. 1995, 60, 45.
(479) Carrano, C. J.; Drechsel, H.; Kaiser, D.; Jung, G.; Matzanke,
B.; Winkelmann, G.; Rochel, N.; Albrecht-Gary, A. M. Inorg.
Chem. 1996, 35, 6429.
(480) Kaliva, M.; Giannadaki, T.; A., S. Inorg. Chem. 2002, 41, 3850.
(481) Zhou, Z.-H.; Zhang, H.; Jiang, Y.-Q.; Lin, D.-H.; Wan, H.-L.; Tsai,
K.-R. Trans. Met. Chem. 1999, 24, 605.
(482) Zhou, Z.-H.; Wan, H.-L.; Hu, S.-Z.; Tsai, K.-R. Inorg. Chim. Acta
1995, 237, 193.
(483) Wright, D. W.; Humiston, P. A.; Orme-Johnson, W. H.; Davis,
W. M. Inorg. Chem. 1995, 34, 4194.
(484) Zhou, Z.-H.; Yan, W.-B.; Wan, H.-L.; Tsai, K.-R.; Wang, J.-Z.;
Hu, S.-Z. J. Chem. Crystallogr. 1995, 25, 807.
(485) Tsaramyrsi, M.; Kaliva, M.; Salifoglou, A.; Raptopoulou, C. P.;
Terzis, A.; Tangoulis, V.; Giapintzakis, J. Inorg. Chem. 2001,
40, 5772.
(486) Velayutham, M.; Varghese, B.; Subramanian, S. Inorg. Chem.
1998, 37, 1336.
(487) Zhou, Z.-H.; Wan, H.-L.; Hu, S.-Z.; Tsai, K.-R. Jiegou Huaxue
1995, 14, 337.
(488) Burojevic, S.; Shweky, I.; Bino, A.; Summers, D. A.; Thompson,
R. C. Inorg. Chim. Acta 1996, 251, 75.
(489) Ehde, P. M.; Andersson, I.; Pettersson, L. Acta Chem. Scand.
1989, 43, 136.
(490) Wright, D. W.; Chang, R. T.; Mandal, S. K.; Armstrong, W. H.;
Orme-Johnson, W. H. J. Biol. Inorg. Chem. 1996, 1, 143.
(491) Dessi, A.; Micera, G.; Sanna, D.; Erre, L. S. J. Inorg. Biochem.
1992, 48, 279.
(492) Luterotti, S. Rev. Anal. Chem. 1992, 11, 195.
(493) Rezaie, B.; Zabeen, R.; Goswami, A. K.; Purohit, D. N. Rev. Anal.
Chem. 1993, 12, 1.
(494) Keller, R. J.; Rush, J. D.; Grover, T. A. J. Inorg. Biochem. 1991,
41, 269.
(495) Brown, D. A.; Boegge, H.; Coogan, R.; Doocey, D.; Kemp, T. J.;
Mueller, A.; Neumann, B. Inorg. Chem. 1996, 35, 1674.
(496) Bell, J. H.; Pratt, R. F. Inorg. Chem. 2002, 41, 2747.
(497) Grigoreva, M. F.; Slesar, N. I.; Tserkovnitskaya, I. A. Zh.
Obshch. Khim. 1982, 52, 1458.
(498) Shukla, J. P.; Tandon, S. G. J. Ind. Chem. Soc. 1972, 49, 83.
(499) Yamaki, R. T.; Paniago, E. B.; Carvalho, S.; Howarth, O. W.;
Kam, W. J. Chem. Soc., Dalton Trans. 1997, 4817.
(500) Yamaki, R. T.; Paniago, E. B.; Carvalho, S.; Lula, I. S. J. Chem.
Soc., Dalton Trans. 1999, 4407.
(501) Chakrabarti, A. K. J. Inst. Chem. (India) 1998, 70, 84.
(502) Fisher, D. C.; Barclay-Peet, S. J.; Balfe, C. A.; Raymond, K. N.
Inorg. Chem. 1989, 28, 4399.
(503) Weidemann, C.; Priebsch, W.; Rehder, D. Chem. Ber. 1989, 122,
235.
(504) Pecoraro, V. L. Inorg. Chim. Acta 1989, 155, 171.
(505) Gibney, B. R. S.; Ann, J.; Pilotek, S.; Kampf, J. W.; Pecoraro, V.
L. Inorg. Chem. 1993, 32, 6008.
(506) Gao, S.; Weng, Z.-Q.; Liu, S.-X. Polyhedron 1998, 17, 3595.
(507) Liu, S.-X.; Gao, S. Inorg. Chim. Acta 1998, 282, 149.
(508) Chen, W.; Gao, S.; Liu, S.-X. Acta Crystallogr. 1999, C55, 531.
Crans et al.
(657) Farahbakhsh, M.; Schmidt, H.; Rehder, D. Chem. Commun.
1998, 2009.
(658) Tsagkalidis, W.; Rodewald, D.; Rehder, D. Inorg. Chem. 1995,
34, 1943.
(659) Farahbakhsh, M.; Nekola, H.; Schmidt, H.; Rehder, D. Chem.
Ber. 1997, 130, 1129.
(660) Tsagkalidis, W.; Rodewald, D.; Rehder, D. J. Chem. Soc., Chem.
Commun. 1995, 165.
(661) Davies, S. C.; Hughes, D. L.; Janas, Z.; Jerzykiewicz, L.;
Richards, R. L.; Sanders, J. R.; Sobota, P. Chem. Commun. 1997,
1261.
(662) Davies, S. C.; Hughes, D. L.; Janas, Z.; Jerzykiewicz, L. B.;
Richards, R. L.; Sanders, J. R.; Silverston, J. E.; Sobota, P. Inorg.
Chem. 2000, 39, 3485.
(663) Kennedy, C.; Dean, D. Mol. Gen. Genet. 1992, 231, 494.
(664) Ruttimann-Johnson, C.; Rangaraj, P.; Shah, V. K.; Ludden, P.
W. J. Biol. Chem. 2001, 276, 4522.
(665) Shilov, A. E.; Denisov, N. T.; Efimov, N. O.; Shuvalov, N.;
Shuvalova, N. I.; Shilova, A. K. Nature 1971, 231, 460.
(666) Nikonova, L. A.; Ovcharenko, A. G.; Efimov, O. N.; Avilov, V.
A.; Shilov, A. E. Kinet. Katal. 1972, 13, 1602.
(667) Denisov, N. T.; Rudshtein, E. I.; Shuvalova, N. I.; Shilova, A.
K.; Shilov, A. E. Dokl. Akad. Nauk 1972, 202, 623.
(668) Ihmels, K.; Rehder, D. Chem. Ber. 1985, 118, 895.
(669) Desmangles, N.; Jenkins, H.; Ruppa, K. B.; Gambarotta, S. Inorg.
Chim. Acta 1996, 250, 1.
(670) Re, N.; Rosi, M.; Sgamellotti, A.; Floriani, C.; Solari, E. Inorg.
Chem. 1994, 33, 4390.
(671) Re, N.; Rosi, M.; Sgamellotti, A.; Floriani, C. Inorg. Chem. 1995,
34, 3410.
(672) Gailus, H.; Woitha, C.; Rehder, D. J. Chem. Soc., Dalton Trans.
1994, 3471.
(673) Deeth, R. J.; Langford, S. A. J. Chem. Soc., Dalton Trans. 1995,
1.
(674) Woitha, C.; Rehder, D. Angew. Chem., Int. Ed. Engl. 1990, 29,
1438.
(675) Rehder, D.; Woitha, C.; Priebsch, W.; Gailus, H. J. Chem. Soc.,
Chem. Commun. 1992, 4, 364.
(676) Berno, P.; Hao, S.; Minhas, R.; Gambarotta, S. J. Am. Chem.
Soc. 1994, 116, 7417.
(677) Spek, A. L.; Edema, J. J. H.; Gambarotta, S. Acta Crystallogr.
1994, C50, 1209.
(678) Clancy, G. P.; Clark, H. C. S.; Clentsmith, G. K. B.; Cloke, F. G.
N.; Hitchcock, P. B. J. Chem. Soc., Dalton Trans. 1999, 3345.
(679) Clentsmith, G. K. B.; Bates, V. M. E.; Hitchcock, P. B.; Cloke,
F. G. N. J. Am. Chem. Soc. 1999, 121, 10444.
(680) Le Floch, C.; Henderson, R. A.; Hitchcock, P. B.; Hughes, D. L.;
Janas, Z.; Richards, R. L.; Sobota, P.; Szafert, S. J. Chem. Soc.,
Dalton Trans. 1996, 2755.
(681) Woitha, C.; Rehder, D. J. Organomet. Chem. 1988, 353, 315.
(682) Tsagkalidis, W.; Woitha, C.; Rehder, D. Inorg. Chim. Acta 1993,
205, 239.
(683) Ferguson, R.; Solari, E.; Floriani, C.; Osella, D.; Ravera, M.; Re,
N.; Chiesi-Villa, N.; Rizzoli, C. J. Am. Chem. Soc. 1997, 119,
10104.
(684) Dilworth, M. J.; Eady, R. R.; Robson, R. L.; Miller, R. W. Nature
1987, 327, 167.
(685) Schneider, K.; Mueller, A.; Krahn, E.; Hagen, W. R.; Wassink,
H.; Knuettel, K.-H. Eur. J. Biochemistry 1995, 34, 666.
(686) Kelly, M.; Postgate, J. R.; Richards, R. L. Biochem. J. 1967, 102,
1C.
(687) Walker, J. V.; Butler, A. Inorg. Chim. Acta 1996, 243, 201.
(688) Eulering, B.; Schmidt, M.; Pinkernell, U.; Karst, U.; Krebs, B.
Angew. Chem., Int. Ed. Engl. 1996, 35, 1973.
(689) Soedjak, H. S.; Butler, A. Biochemistry 1990, 29, 7974.
(690) Arends, I. W. C. E.; Vos, M.; Sheldon, R. A. ACS Symp. Ser.
1998, 711, 146.
(691) Rehder, D.; Bashirpoor, M.; Jantzen, S.; Schmidt, H.; Farahbakhsh, M.; Nekola, H. ACS Symp. Ser. 1998, 711, 60.
(692) Bolm, C.; Bienewald, F. Angew. Chem., Int. Ed. Engl. 1996, 34,
2640.
(693) Everett, R. R.; Butler, A. Inorg. Chem. 1989, 28, 393.
(694) Doctrow, S. R.; Huffman, K.; Bucay Marcus, C.; Tocco, G.;
Malfroy, E.; Adinolfi, C. A.; Kruk, H.; Baker, K.; Lazarowych,
N.; Mascarenhas, J.; Malfroy, B. J. Med. Chem. 2002, 45, 4549.
(695) Burton, J. D. Nature 1966, 212, 976.
(696) Premovic, P. I.; Nikolic, N. D.; Pavlovic, M. S.; Jovanovic, L. S.;
Premovic, M. P. Geol. Soc. Spec. Publ. 1997, 125, 273.
(697) Afzal, D.; Baughman, R.; James, A.; Westmeyer, M. Supramol.
Chem. 1996, 6, 395.
(698) Drew, M. G. B.; Mitchell, P. C. H.; Scott, C. E. Inorg. Chim. Acta
1984, 82, 63.
(699) Poncet, J. L.; Barbe, J. M.; Guilard, R.; Oumous, H.; Lecomte,
C.; Protas, J. J. Chem. Soc., Chem. Commun. 1982, 1421.
(700) Wang, X.; Gray, S. D.; Chen, J.; Woo, L. K. Inorg. Chem. 1998,
37, 5.
(701) Berreau, L. M.; Hays, J. A.; Young, V. G., Jr.; Woo, L. K. Inorg.
Chem. 1994, 33, 105.
Crans et al.
(797) Nagaoka, M. H.; Yamazaki, T.; Maitani, T. Biochem. Biophys.
Res. Commun. 2002, 296, 1207.
(798) Ghio, A. J.; Carter, J. D.; Samet, J. M.; Reed, W.; Quay, J.;
Dailey, L. A.; Richards, J. H.; Devlin, R. B. Am. J. Physiol. 1998,
274, L728.
(799) Hirsch, S.; Miskimins, R.; Miskimins, W. K. Recept. Signal
Transduction 1996, 6, 121.
(800) Antipov, A. N.; Lyalikova, N. N.; Khijniak, T. V.; Lvov, N. P.
FEBS Lett. 1998, 441, 257.
(801) Antipov, A. N.; Sorokin, D. Y.; LVov, N. P.; Kuenen, J. G.
Biochem. J. 2003, 369, 185.
CR020607T