You are on page 1of 10

View Article Online / Journal Homepage / Table of Contents for this issue

Metallomics Dynamic Article Links

Cite this: Metallomics, 2011, 3, 1153–1162

www.rsc.org/metallomics CRITICAL REVIEW


Mechanisms of nickel toxicity in microorganismsw
Published on 28 July 2011. Downloaded by University of Florida Libraries on 27/12/2017 07:28:16.

Lee Macombera and Robert P. Hausinger*ab


Received 10th June 2011, Accepted 14th July 2011
DOI: 10.1039/c1mt00063b

Nickel has long been known to be an important human toxicant, including having the ability to
form carcinomas, but until recently nickel was believed to be an issue only to microorganisms
living in nickel-rich serpentine soils or areas contaminated by industrial pollution. This
assumption was overturned by the discovery of a nickel defense system (RcnR/RcnA) found in
microorganisms that live in a wide range of environmental niches, suggesting that nickel
homeostasis is a general biological concern. To date, the mechanisms of nickel toxicity in
microorganisms and higher eukaryotes are poorly understood. In this review, we summarize
nickel homeostasis processes used by microorganisms and highlight in vivo and in vitro effects of
exposure to elevated concentrations of nickel. On the basis of this evidence we propose four
mechanisms of nickel toxicity: (1) nickel replaces the essential metal of metalloproteins, (2) nickel
binds to catalytic residues of non-metalloenzymes; (3) nickel binds outside the catalytic site of an
enzyme to inhibit allosterically and (4) nickel indirectly causes oxidative stress.

1. Introduction is typically found in the Ni(0) or Ni(II) state due to the stability
of these species in water.3 Nickel toxicity to humans has
Nickel concentrations vary widely in different environmental received intensive attention due to nickel’s links to cancer.4–7
niches, with low nanomolar [i.e., B0.1 ng g1 (ppb)] to low Similarly, nickel toxicity to plants is well documented.8
micromolar [i.e., B0.1 mg g1 (ppm)] levels in many aquatic Until recently, nickel toxicosis in microorganisms was believed
systems, 10–40 ppm in most soils, and several 100 ppm in to be a problem limited to cells exposed to industrial pollution
serpentine soils.1 In addition, nickel is a common industrial or found in naturally-occurring nickel-contaminated soils.9
pollutant with concentrations in wastewater around industrial This assumption was overturned, however, when a
parks sometimes reaching the low millimolar range.2 This metal nickel defense system (RcnA) was characterized in the meso-
phile Escherichia coli.10 Furthermore, putative homologues of
a
Department of Microbiology and Molecular Genetics, RcnA were detected in archaea and throughout bacteria.10
Michigan State University, East Lansing, Michigan 48824-4320, This discovery suggests that excessive nickel is a routine
USA. E-mail: hausinge@msu.edu physiological concern of microorganisms. For example,
b
Department of Biochemistry and Molecular Biology, Michigan State commensal bacteria in the human gut are exposed to
University, East Lansing, Michigan 48824-1319, USA
w This article is published as part of a themed issue on Metal Toxicity, 150–900 mg of nickel per day as a constituent of the diet.11
Guest Edited by Gregor Grass and Christopher Rensing. In this review we summarize what is known about Ni(II)

Lee Macomber received his Robert P. Hausinger received


PhD with James A. Imlay at his PhD at the University of
the University of Illinois Minnesota in 1982. After post-
Urbana-Champaign in 2009. doctoral training at M.I.T.
He is undergoing postdoctoral (1982–1984) he joined the
training with Robert P. faculty at Michigan State
Hausinger at Michigan State University where he is a
University, and is studying nickel Professor of Microbiology &
homeostasis in Escherichia coli Molecular Genetics and
and Bacillus subtilis. Biochemistry & Molecular
Biology. His research focuses
on metallocenter biosynthesis
and the catalytic mechanisms
of metalloenzymes.
Lee Macomber Robert P. Hausinger

This journal is c The Royal Society of Chemistry 2011 Metallomics, 2011, 3, 1153–1162 1153
View Article Online

homeostasis and toxicity in microorganisms and we discuss eubacteria and archaea utilizes an ATP-binding cassette
potential mechanisms of nickel toxicosis. (ABC) system, such as NikABCDE, whereas the most
common mechanism in eukaryotes (also widely distributed
2. Nickel homeostasis in microorganisms in bacteria) involves members of the nickel/cobalt transporter
(NiCoT) family. In addition, nonspecific transporters of nickel
Nickel is used as a cofactor by several well-characterized are used by some microorganisms to facilitate transfer of the
microbial enzymes including urease, [NiFe] hydrogenase, metal across the cytoplasmic membrane. Finally, bacteria with
Ni-superoxide dismutase, carbon monoxide dehydrogenase, an outer membrane have, in some cases, evolved mechanisms to
acetyl CoA synthase/decarbonylase, acireductone dioxygenase,
Published on 28 July 2011. Downloaded by University of Florida Libraries on 27/12/2017 07:28:16.

expedite nickel transport across that barrier. These topics, briefly


and methyl coenzyme M reductase, as well as some forms of summarized below, have been reviewed previously12,13,38,39 and
glyoxalase I.12–15 Additional nickel-dependent enzyme are illustrated in the lower portion of Fig. 1.
activities are reported, and include glycerol-1-phosphate NikABCDE was first characterized in E. coli and shown to
dehydrogenase from Bacillus pasteurii16 and quercetinase from be a part of the ABC transporter subfamily 2.40 NikB and
a Streptomyces species.17 The metallocenter of nickel enzymes NikC are transmembrane proteins that form a nickel pore.40
typically is coordinated by histidine and/or cysteine residues, NikD and NikE are proteins that bind to and hydrolyze
with additional contributions from aspartate, glutamate and, ATP.40 NikA is a periplasmic protein that binds one nickel
in the case of urease, a lysine carbamate. In addition, some per protein along with a naturally-occurring organic metallo-
proteins bind nickel using the amino terminal amine or back- phore; the structure of this nickel chelate remains unknown
bone amide groups.12–15 but is thought to be a tricarboxylate.41 Nik homologues,
In organisms with nickel-containing proteins, mechanisms sometimes known as Cbi/NikMNQO and often lacking the
have evolved to take up the metal. At the same time, nickel periplasmic component, have been found in many bacteria. A
efflux or other detoxification systems have evolved to over- few examples of other Nik systems include those in Brucella suis,42
come problems of excessive concentrations of this metal. Vibrio parahemolyticus,43 Helicobacter hepaticus,44 Yersinia
Accompanying these uptake and resistance mechanisms, sp.,45 and Staphylococcus aureus.46 The NikABCDE import
microorganisms often contain proteins for sensing and respond- system is transcriptionally regulated by NikR,25 whose gene is
ing to nickel. Finally, some microbes appear to employ nickel in turn regulated by FNR.47
storage proteins.15 Together these processes provide mechan- The founding member of the NiCoT family is HoxN, a
isms for nickel homeostasis in microorganisms.12,13,18–22 high-affinity nickel permease found in Cupriavidus necator
Microbes generally concentrate metals from their growth (formerly Alcaligenes eutrophus) and required for full urease
medium,23,24 and this is true also for nickel. For example, and hydrogenase activities.48–50 This family can be divided
E. coli cells grown on M9 minimal medium (containing B1 mM into those specific to nickel, those capable of transporting both
nickel) contained B30 mM of this metal (unpublished nickel and cobalt, and those with a preference for cobalt.39
observation). Whereas the total concentration of nickel in a NiCoT proteins generally exhibit a low Km (10–20 nM) and a
microorganism can be ascertained, this value does not equate low transport capacity for their metal.50,51 NiCoT representa-
to the concentration of the free metal in the cell. tives have been identified in a wide range of bacteria, including
Rhodococcus rhodochrous,52,53 Bradyrhizobium japonicum,54
2.1 Nickel metalloregulatory proteins
H. pylori,55 and Mycobacterium tuberculosis.56 Family members
Several DNA-binding proteins undergo conformational additionally are prevalent in eukaryotic microorganisms,39
changes upon binding nickel, leading to changes in transcrip- as first demonstrated for Schizosaccharomyces pombe.57
tion of genes encoding metal transport and nickel proteins.
NikR is the best-studied nickel-dependent regulatory protein.25
Tetrameric NikR from E. coli binds four nickel at high-affinity
sites,26,27 with additional sites for binding potassium and
(with lower affinity) additional nickel,28 and the nickel-bound
protein serves as a repressor by binding to specific operator
sequences.29 NikR proteins of other organisms exhibit
differences in metal binding and regulatory properties.30–33
Working in the opposite manner is RcnR, an E. coli protein
that releases from the rcnA promoter DNA in the presence of
nickel, thus allowing transcription to take place.34 Studies
on the ArsR-SmtB family of metal-sensing transcriptional
repressors in Mycobacterium tuberculosis discovered two
members (NmtR and KmtR) that sense nickel and cobalt.35,36
Some of the actions of these regulatory proteins are high-
lighted in the following sections.

2.2 Nickel import


According to comparative genomic analysis,37 the most wide- Fig. 1 A representation of selected nickel trafficking pathways in
spread mechanism for high-affinity uptake of nickel in bacteria.

1154 Metallomics, 2011, 3, 1153–1162 This journal is c The Royal Society of Chemistry 2011
View Article Online

The energy source required to import nickel by NiCoT Finally, C. metallidurans also contains a chromosomally-
proteins remains unknown.13 encoded DmeF efflux pump that confers resistance to iron,
Nonspecific nickel import into the cytoplasm also occurs. zinc, cobalt, cadmium, and nickel.72 The corresponding gene is
For example, nickel can enter E. coli through the magnesium constitutively expressed and is not induced by any of the listed
importer CorA.58 Deletion of E. coli genes encoding both Nik metals, but additional findings led to the conclusion that
and CorA still left the strains nickel sensitive,40 therefore this DmeF plays a central role in cobalt homeostasis.72
metal must be capable of entering the cytoplasm through other The RcnA efflux pump was first identified in E. coli and
channels. Entry of nickel through CorA and other adventi- shown to pump nickel and cobalt out of the cytoplasm.10
Published on 28 July 2011. Downloaded by University of Florida Libraries on 27/12/2017 07:28:16.

tious transporters likely serves as the route of nickel entry Putative homologues exist in alpha, beta, and gamma proteo-
under high environmental nickel concentrations. bacteria, cyanobacteria, and archaea.10 Expression of rcnA is
Nickel must pass an additional barrier in eubacteria regulated by the nickel/cobalt responsive repressor, RcnR. As
containing an outer membrane. The metal initially was expected, the relative nickel-binding affinities of RcnR
thought to passively cross through porins; however, H. pylori (Kapp o 26 nM) and NikR (Kd r 1 pM)19 determine the
urease activity was shown to be inhibited by knocking out the differential transport activities of RcnA and NikABCDE, and
TonB-dependent transporters (TBDTs) FecA3 and FrpB4.59,60 thus poise the internal nickel concentration of the cell.19 The
Putative TBDTs of nickel have since been found in several discovery of the chromosomally-encoded RcnA in the
other organisms.61,62 mesophile E. coli suggests that environmental nickel is a
In summary, several distinct types of mechanisms are used concern to many microorganisms under typical conditions.
to import nickel into microbial cells. This metal has a bene- Additional representative exporters of nickel include a
ficial role when incorporated into nickel-dependent enzymes, putative metal efflux P1 type ATPase (nmtA), repressed by
but it can be toxic when present in excessive amounts. NmtR, and a putative metal efflux pump (cdf) of the cation-
diffusion-facilitator family, repressed by KmtR.35,36
2.3 Nickel efflux
2.4 Other nickel resistance mechanisms
Microbes are subject to their environmental conditions and, as
such, they are exquisitely vulnerable to changes in metal Although nickel efflux is widely used by cells to protect against
concentrations. Compounding this issue is their tendency to elevated concentrations of this metal, several other mechan-
concentrate metals.23,24 For these reasons microorganisms isms are utilized by selected microorganisms. E. coli exhibits a
have evolved ways to protect themselves from metal overload. chemotactic response to environmental nickel concentrations,
A common microbial response to elevated concentrations of a with nickel acting as a chemotactic repellent.73–75 This
toxic metal is to synthesize a specific efflux system, thus response is dependent on the methyl-accepting chemotaxis
reducing the internal concentration of that metal species.63 protein Tar and is independent of NikA.75 Nickel-resistant
Nickel efflux pumps are best characterized in organisms strains of Saccharomyces cerevisiae have long been known to
exhibiting hyper-resistance to this metal, typically isolated sequester the metal in a vacuolar fraction along with high
from nickel-rich, polluted or naturally-occurring soils. Two concentrations of the amino acid histidine.76,77 Genetic studies
examples of nickel-resistant organisms obtained from metal- revealed that histidine auxotrophs exhibit enhanced sensitivity
contaminated industrial sites are Cupriavidus (formerly to nickel, along with several other metals, and this work
Wautersia, Ralstonia, or Alcaligenes) metallidurans and showed that strains containing a defective vacuolar proton
Alcaligenes (or Achromobacter) xylosoxidans.64,65 Nickel efflux ATPase (thus unable to acidify the vacuole) display greater
pumps also are present in non-extremophiles, as exemplified nickel toxicity.78 In another yeast study, genetic knockouts of
by E. coli10 and H. pylori.66 Below, we summarize features of iron import sensitized the cells to nickel.79 More generally,
nickel efflux (see the upper half of Fig. 1) in these micro- genome-wide analyses of S. cerevisiae identified 149 genes
organisms as representatives of many other species. that, when deleted, led to increased nickel sensitivity consistent
Work in C. metallidurans, A. xylosoxidans, and H. pylori with having a role in nickel resistance.80 Some of these genes
identified several nickel-resistance determinants. The plasmid- relate to proton transporting ATPases, but cation transpor-
borne gene clusters cnrCBA (cobalt and nickel resistance)67 ters, siderophore-iron transporters, diphthamide biosynthesis,
and nccCBA (nickel-cobalt-cadmium resistance)68 and the and other processes are included. Sulfate-reducing bacteria
chromosomal cznCBA (cadmium, zinc, and nickel resistance)66 produce copious amounts of sulfide that can abrogate the
encode members of the resistance nodulation division (RND) toxicity of the metal. For example, a nickel-resistant strain of
family containing inner membrane, periplasm, and outer Desulfotomaculum cultured with high concentrations of this
membrane components and driven by the proton gradient. metal produces a dark-brown, soluble, nickel-sulfide product
These proteins are believed to pump periplasmic metals out as a means to reduce the concentration of the free ion.81
across the outer membrane, and thus are not found in bacteria Pseudomonas putida S4 accumulates nickel in its periplasm,
lacking outer membranes.69 Another plasmid-encoded possibly in complex with an 18 kDa protein, as a mechanism
protein, NreB, is part of the major facilitator superfamily of resistance to this metal.82 By contrast, a strain of
(MFS) and likely utilizes the proton motive force to pump Pseudomonas aeruginosa accumulates intracellular nickel
nickel out of the cytoplasm.70 Expression of nreB in which, according to energy-dispersive X-ray analysis, is in
C. metallidurans or E. coli increases cellular resistance to nickel the form of metallic nickel.83 The reduction of Ni(II) to
challenge. Homologues of nreB exist in several other bacteria.63,71 elemental nickel also is observed in Thiocapsa rosepersicina

This journal is c The Royal Society of Chemistry 2011 Metallomics, 2011, 3, 1153–1162 1155
View Article Online

when grown on hydrogen gas, and this reduction is linked to four general hypotheses: (1) nickel replaces the essential metal
greater nickel tolerance.84 of metalloproteins, (2) nickel binds to catalytic residues of
The distribution of nickel defense systems in hyper-resistant non-metal enzymes, (3) nickel binds outside the catalytic site
as well as more common microorganisms including E. coli of an enzyme to inhibit allosterically and (4) nickel leads to
suggests that excessive nickel is a widespread physiological oxidative stress that can affect proteins, DNA, or lipids. These
concern. This leads to the question: What do the nickel defense generalized toxicity mechanisms are examined in greater detail
systems protect? The following section focuses on potential below with examples of binding at secondary sites folded in
cellular targets of nickel toxicity in microorganisms. with the inhibition of metalloenzymes and non-metal enzymes.
Published on 28 July 2011. Downloaded by University of Florida Libraries on 27/12/2017 07:28:16.

3.1 Nickel toxicity involving metalloproteins


3. Nickel toxicity
Metals participate in a variety of essential biological functions,
As illustrated by the plethora of review articles on nickel including metalloregulation, structural stabilization, electron
toxicity toward humans, other animals, and plants (e.g., transfer, substrate/cofactor coordination, and catalysis. The
ref. 4–8), much is known about how this metal harms the cells atomic sizes, ligand affinities, preferred coordination geome-
of higher eukaryotes. Much of the early literature focused on tries, redox states, and available metal concentrations dictate,
the toxic and carcinogenic effects of particulate nickel com- in part, the identity of the bound metals, with accessory
pounds on ore miners, with additional studies relating to the proteins also controlling the metal speciation in selected
toxicity of nickel carbonyl (used in refining) and immunological enzymes.89 Although cells attempt to maintain precise intra-
effects in nickel dermatitis. The mechanisms of nickel damage cellular metal concentrations, this homeostasis can fail when
to humans had long emphasized oxidative reactions involving excess nickel is present and lead to adventitious displacement
lipids, proteins, and DNA, but nickel also was known to bind of the cognate metal. Replacement of the normal metal by
to various biomolecules and change their properties.4–7 Only nickel in a metalloregulatory protein could reasonably inter-
recently has nickel been shown to have profound effects on fere with its function. In contrast, substitution of nickel into a
DNA repair and DNA or histone methylation, in the latter structural site is unlikely to exhibit severe repercussions.
case resulting in epigenetic effects.85–87 In contrast to this Incorporation of this metal into an electron transfer site that
wealth of information, the mechanisms of nickel toxicity in normally contains iron or copper likely will inhibit function
microorganisms have been understudied, despite the widely due to the stable nature of the Ni(II) state. Below, we focus on
demonstrated toxic effect of this metal.9 a few illustrative examples where nickel either substitutes for a
Several strategies can be used to establish the cellular targets catalytic metal ion or binds at a second site on a metallo-
of nickel toxicosis. At the organismal-level, clues may be enzyme to inhibit catalysis (Fig. 2 and Table 1).
gained by examining the nickel sensitivity for cells grown
under varied conditions (e.g., use of different nutrients, aerobic 3.1.1 Nickel inhibition of iron metalloenzymes. A clear
vs. anaerobic conditions, treatment with other stresses, etc.), connection exists between nickel toxicity and iron metabolism.
by transcriptomic or proteomic comparisons to identify genes In E. coli, Fur was found to regulate rcnA in an iron-
or proteins synthesized during nickel stress, and by mutational dependent manner.90 Also, iron supplementation induced
analysis to identify genes that, when mutated, result in rcnR transcription in a Fur-independent manner.90 Further-
enhanced resistance to the metal. An example of the proteomic more, the Fur-regulated gene yqjH of E. coli, shown to be a
approach is provided by analysis of nickel-stressed ferric reductase that plays a role in iron acquisition, is also
Pseudomonas putida.88 A range of proteins with altered regulated by nickel through YqjI.91,92 The authors demon-
expression was identified including several up-regulated strated that yqjH mutants are sensitive to micromolar levels of
proteins involved in antioxidant defense. An extensive deletion nickel under conditions that require significant production of
mutant screen of S. cerevisiae nicely demonstrates how that the [FeS]-cluster enzyme 6-phosphogluconate dehydratase for
approach can be exploited.80 Enhanced resistance was
associated with deletions of 119 genes, including those causing
derepression of glucose-repressed genes. In addition to the
above organismal approaches, more targeted studies to define
the sites of nickel toxicity focus on direct biochemical assays in
the presence and absence of the metal ion. For example, the
effects of nickel on DNA or lipid modifications have been
examined. In addition, enzyme activities can be measured
using intact cells, cell-free extracts, partially enriched samples,
or purified enzymes with the results used to identify potential
targets of nickel inhibition. Several examples of nickel-
inhibited enzymes are described below. It is important to
emphasize, however, that nickel inhibition of an enzyme Fig. 2 A schematic representation of potential mechanisms of nickel
activity measured in vitro must be verified for its in vivo toxicity. Mechanisms are divided into three categories: direct mecha-
relevance—a situation that is seldom realized. nisms with multiple lines of evidence (black lines), direct mechanisms
Observations from nickel-toxified cells and measurements of with few lines of evidence (dashed lines), and indirect mechanisms
nickel’s in vitro effects on biomolecules can be synthesized into (blue lines).

1156 Metallomics, 2011, 3, 1153–1162 This journal is c The Royal Society of Chemistry 2011
View Article Online

Table 1 Representative examples of microbial enzymes inhibited by nickel

Active Site Enzyme Organism [Ni(II)] mM Inactivation (%) Ref.


Iron taurine/aKG dioxygenase E. coli 32 50 96
thymidine 2 0 hydroxylase N. crassa 250 Z 95 97
protocatechuate 3,4-dioxygenase R. leguminosarum 10 85 99
catechol 1,2-dioxygenase R. leguminosarum 100 50 100
ATP:cob(I)alamin adenosyltransferase S. enterica 100 50 102
atrazine chlorohydrolase Pseudomonas sp. 200 60 105
Zinc metallocarboxypeptidase (MCP-1) T. cruzi 10 54 106
Published on 28 July 2011. Downloaded by University of Florida Libraries on 27/12/2017 07:28:16.

aminopeptidase S. septatus 100 47 107


aminopeptidase S. griseus 100 60 107
tripeptidase P. pentosaceus 100 94 110
prenyl protease (Rce1p) S. cerevisiae 87 48 111
Copper nitrous oxide reductase R. sphaeroides 100 50 112
Cobalta cobaltochelatase P. denitrificans 3 50 113
Magnesium DNA polymerase I (Klenow fragment) E. coli 370 50 114
DNA polymerase T4 bacteriophage 240 50 114
DNA polymerase T7 bacteriophage 780 50 114
Non-metal N-carbamoyl D-amino acid amidohydrolase Agrobacterium radiobacter 0.1 NSb 117
pyroglutamyl peptidase I L. major 100 48 120
mercury(II) reductase A. chroococcum 100 62 122
cytosine-5 methyltransferase E. coli 1 Z 80 123
uracil-5 methyltransferase E. coli 1 Z 80 123
a b
Cobalt is a substrate for this enzyme, not a required cofactor. NS, 0.1 mM was said to be inhibitory, but the extent of inhibition was not stated.

growth. These results suggest that E. coli may need increased


iron uptake during nickel overload,92 a situation that is likely
to apply to many more organisms. A requirement for
increased iron when cells are stressed with nickel is compatible
with the mis-incorporation of nickel into iron metalloenzymes.
As illustrated below, many iron-containing enzymes are
inhibited by nickel, with some shown to have nickel replacing
the active metal.
Many iron- and a-ketoglutarate (aKG)-dependent dioxy-
genases are inhibited by nickel, likely via direct displacement
of the active site iron (Fig. 3). These enzymes possess a
mononuclear, nonheme ferrous ion at the active site and
couple the oxidative decarboxylation of aKG to the oxidation
of their diverse primary substrates.93–95 One of the better
Fig. 3 Nickel damage involving substitution of the functional metal
studied examples is taurine/aKG dioxygenase (TauD), an
in a metalloenzyme. The structure is shown in cartoon mode for one
E. coli enzyme that releases sulfite from aminoethanesulfonate
subunit of the iron(II)- and a-ketoglutarate-dependent E. coli protein
as part of a sulfur assimilation pathway. In vitro experiments TauD (PDB access code 1os7).172 Nickel (green sphere) is known to
starting with the iron-loaded enzyme measured an IC50 for inhibit this protein by a slow-binding process involving competition
nickel of 32 mM, whereas the IC50 decreased to 0.7 mM Ni(II) if with iron (brown sphere) for its amino acid ligands (shown as sticks
iron was added after nickel exposure.96 These results imply with orange carbons) and a-ketoglutarate (sticks with yellow carbons).
that taurine/aKG dioxygenase will be more susceptible to Figure prepared with PYMOL.
nickel during iron limitation or immediately after synthesis.
Notably, nickel was shown to exhibit slow-binding inhibition
kinetics with an initial Ki of 166 mM and a maximum inactivation Inhibition of these chromatin modification enzymes may
rate constant of 0.078 s1; thus, nickel is not a potent inhibitor explain the nickel-dependent mutagenic and epigenetic pheno-
of TauD. Another iron/aKG-dependent dioxygenase is the types observed in mammals.6,85,98 Nearly all members of this
fungal enzyme thymidine 2 0 -hydroxylase (pyrimidine- enzyme family contain an iron-binding motif (comprised of
deoxynucleoside 2 0 -dioxygenase) from Neurospora crassa His-X-Asp/Glu-Xn-His) that binds iron using only three
which lost Z 95% of its in vitro ability to hydroxylate residues and is, thus, likely to compete for nickel. Significantly,
pyrimidines upon exposure to 250 mM Ni(II).97 Inhibition of structures of several family members, crystallized in the
the Fe/aKG-dependent dioxygenases in microorganisms presence of nickel to avoid iron-based redox chemistry, show
parallels the findings in mammalian systems where nickel the inhibitory metal is bound to the iron-binding motif
inhibits histone demethylases (JMJD1A, JMJD2, and JMJD2E), (e.g., PDB ID codes 2ox0, 2w2i, 2wwu, 2xlm, 2xxz, 3k2o,
oxygen-sensing prolyl 4-hydroxylases (PHD2, EGLN1), 3ai6, and 3ms5).
and DNA repair nucleotide demethylases (ABH2 and Several other ferrous ion-dependent dioxygenases also are
ABH3) by replacement of the catalytic iron atom.85–87 inhibited in vitro by nickel, presumably due to the relatively

This journal is c The Royal Society of Chemistry 2011 Metallomics, 2011, 3, 1153–1162 1157
View Article Online

redox-stable inhibitor replacing the active metal. For example, 3.1.3 Nickel inhibition of other metalloenzymes. An exhaus-
protocatechuate 3,4-dioxygenase from Rhizobium leguminosarum tive search likely would uncover a large number of other
(trifolii) lost 72% of its activity after exposure to 10 mM metalloenzymes that are inhibited in vitro by nickel, but in
Ni(II):EDTA.99 This enzyme cleaves the aromatic ring of few cases are any mechanistic details available. Three
protocatechuate by adding dioxygen to form b-carboxy-cis,cis- examples are illustrated here. The copper-containing enzyme
muconate, thus playing an important role in aromatic catabolism. nitrous oxide reductase from Rhodobacter sphaeroides112 is
In an analogous manner, catechol 1,2-dioxygenase cleaves inhibited by nickel. Interestingly, zinc inhibition of the enzyme
catechol to cis,cis-muconate. When exposed to 10 mM did not lead to loss of the copper112 which could mean that
Published on 28 July 2011. Downloaded by University of Florida Libraries on 27/12/2017 07:28:16.

Ni(II):EDTA this enzyme, also from R. leguminosarum, nickel inhibition also may be independent of copper loss. The
similarly lost 85% of its activity.100 second example concerns a case where the enzyme uses metal
In addition to ferrous ion-dependent dioxygenases, many as the substrate rather than for catalysis. Cobaltochelatase
other iron-dependent enzymes have been shown to be sensitive from Pseudomonas denitrificans inserts cobalt into hydrogeno-
to nickel. As one example, ATP:cob(I)alamin adenosyltrans- byrinic acid a,c-diamine during vitamin B12 biosynthesis, and
ferase (EutT) from Salmonella enterica is thought to be an this activity was shown to be blocked by nickel.113 The exact
iron-containing enzyme101–103 that catalyzes the adenylation mechanism of cobalt insertion has yet to be determined. It is
of cobalamin (vitamin B12) to adenosylcobalamin (coenzyme B12) conceivable that nickel competes with cobalt for binding to the
for use in ethanolamine catabolism.101–104 Purified EutT lost enzyme or that nickel binds to and thus sequesters the required
50% of its activity upon exposure to 100 mM Ni(II).102 Of thiol. Finally, nickel has variable effects on the magnesium-
interest, nickel concentrations higher than 100 mM did not dependent DNA polymerases.114 For example, nickel substi-
decrease EutT activity below 50%102 implying either that tutes for the usual metal in E. coli DNA polymerase I, is a
nickel binds to an allosteric site, rather than displacing the weak inhibitor of the bacteriophage T4 enzyme, and is a more
catalytic iron, or that EutT:Ni catalyzes the adenylation of effective inhibitor of the enzyme from bacteriophage T7.
cobalamin at a reduced rate compared to the iron-bound
species. Atrazine chlorohydrolase (AtzA) from a Pseudomonas sp. 3.2 Nickel toxicity involving non-metal enzymes
is another putative iron enzyme that is sensitive to nickel
inhibition.105 This enzyme hydrolytically dechlorinates the Nickel also is known to inhibit a wide selection of enzymes
herbicide atrazine to form hydroxyatrazine. Exposure of that do not require a metal for catalysis. In general terms, such
isolated enzyme to 200 mM Ni(II) led to the loss of 60% AtzA inhibition could occur by nickel binding to active site residues
activity.105 The authors determined that Zn(II) and Cu(II) also (Fig. 4), thus blocking catalysis, or, as mentioned above, by
inhibit the enzyme, in these cases without displacing the active the metal binding at a secondary site to influence activity
site metal, so it is possible that nickel inhibits this enzyme by allosterically (Fig. 2). In most cases, the mechanism of inhibi-
binding at an allosteric site.105 tion is unknown.
Several enzymes shown to be inhibited by nickel are known
3.1.2 Nickel inhibition of zinc metalloenzymes. Several zinc to contain catalytic cysteine residues. For example, N-carbamoyl
D-amino acid amidohydrolase from Agrobacterium uses a
metallopeptidases have been shown to be susceptible to
micromolar levels of nickel. The Trypanosoma cruzi metallo- Cys/Glu/Lys catalytic triad to produce D-amino acids115,116
carboxypeptidase 1 (TcMCP-1), which cleaves the carboxyl- which are important for the production of b-lactam antibiotics.
terminal amino acid from peptides and proteins, is one Wang et al.116 found that the enzyme additionally contains
example. Purified TcMCP-1 lost 54% of its activity upon three histidine residues important for enzyme function. This
incubation with 10 mM Ni(II).106 This enzyme utilizes two enzyme is sensitive to nickel-dependent inhibition at concen-
histidines (H267 and H271) and a glutamate (E296) to trations as low as 100 nM,117 though its mode of binding the
coordinate the active site zinc,106 and these residues are likely
to function in nickel binding. Similarly, aminopeptidases
from Streptomyces septatus (SSAP) and Streptomyces griseus
(SGAP) lost 47% and 60% of their activities, respectively,
after incubation with 100 mM Ni(II).107 SSAP and SGAP are
members of a metallopeptidase family containing dinuclear
zinc sites.107–109 One zinc atom is coordinated by a histidine
and an aspartate, the other is coordinated by a second
histidine and glutamate, and the two zinc atoms are bridged
by an aspartate.108 A tripeptidase from Pediococcus pentosaceus,
also suggested to contain a dinuclear zinc site, was found to
lose 94% of its activity after exposure to 100 mM Ni(II).110
Fig. 4 Nickel damage by binding to a non-metal enzyme. The
Prenyl protease Rce1p, a putative zinc metallopeptidase
structure in cartoon mode is shown for N-carbamoyl D-amino acid
from S. cerevisiae, also is sensitive to nickel with mid- amidohydrolase from Agrobacterium radiobacter (PDB access code
micromolar levels (87 mM) leading to 48% inhibition of the 1fo6).116 Nickel (green sphere) is likely to inhibit the protein by
enzyme.111 The mechanism by which nickel inhibits these binding to some combination of three active site residues (depicted
enzymes was not determined, but competition between zinc as sticks with orange carbon atoms) and three nearby histidine
and nickel for enzyme binding is likely. residues (sticks with red carbons). Figure prepared with PYMOL.

1158 Metallomics, 2011, 3, 1153–1162 This journal is c The Royal Society of Chemistry 2011
View Article Online

inhibitory metal is unknown. Pyroglutamyl peptidase I, a of a ligand to act as a prooxidant depends on a low molar ratio
member of the C15 family of cysteine peptidases, functions to nickel which is more likely to provide an open coordination
to hydrolyze N-terminal L-pyroglutamate residues which may site for the binding of the oxidant.148
play a role in protein metabolism and defense against anti-
biotic peptides.118,119 Exposure of the enzyme from Leishmania 3.3.2 In vivo studies. While there is a dearth of microbial
major to 100 mM Ni(II) causes the loss of 48% of its activity.120 studies pertaining to the role of nickel in cellular oxidative
Mercury(II) reductase plays a role in reduction of Hg(II) to stress, some evidence suggests that such stress may occur
Hg(0), thus detoxifying the solution by converting the metal during nickel overload. In proteomics studies that investigated
the effect of nickel toxicity on Burkholderia vietnamiensis151
Published on 28 July 2011. Downloaded by University of Florida Libraries on 27/12/2017 07:28:16.

ion to a volatile species. The enzyme from E. coli (MerA)


contains four cysteine residues121 which might relate to why and P. putida88 the number of oxidative stress-associated
the protein from A. chroococcum is sensitive to 100 mM Ni(II), defense proteins was found to increase. For example,
causing a loss of 62% activity.122 The tRNA-modifying B. vietnamiensis cells stressed with 3.4 mM Ni(II) exhibited
enzymes cytosine-5 methyltransferase and uracil-5 methyl- increases in the level of superoxide dismutase (SOD), a
transferase from E. coli strain W are sensitive to nickel.123 superoxide scavenger.151 Similarly, nickel-toxified P. putida
These enzymes contain catalytic cysteine residues which may produced increases in SOD as well as thioredoxin, ferredoxin
explain why activity for both enzymes decreases by Z 80% NADP reductase, a thiol-specific antioxidant, GTP-binding
upon incubation with 1 mM Ni(II).123–125 protein TypA, and a 2-oxo-acid dehydrogenase.88 Further-
Many other enzymes lacking Cys at their active sites are more, elimination of cytoplasmic SOD sensitized E. coli cells
inhibited by nickel, but the inhibitory mechanisms are generally to this metal.152 Nickel (100 mM) completely inhibited the
unknown. Two examples are cited. Arabinose isomerase, an growth of a sodA sodB mutant on complex medium while
enzyme utilizing catalytic histidine and glutamate residues,126 single mutants were as resistant to nickel as wild-type cells.152
loses 60% of its activity when incubated with 10 mM Ni(II).127 These data suggest that superoxide may be produced in
Uricase (urate oxidase) from Bacillus fastidiosus exposed to nickel-toxified cells. In addition to increased SOD levels,
100 mM Ni(II) loses 81% of its activity;128 this enzyme may nickel-overload led to an increase in two hydrogen peroxide
employ a histidine-threonine-lysine triad for activity.129,130 scavengers, catalase153 and peroxidase.153,154
One of the hallmarks of oxidative stress is damage to DNA.
3.3 Nickel toxicity via oxidative stress Nickel-dependent DNA damage has been documented in
Oxidative stress has long been discussed as a contributing higher eukaryotes,4–7 but the evidence in microorganisms is
factor to nickel carcinogenesis in humans.4–7 As such, there is murky. Nickel was not found to cause reverse mutations in
much interest in studying the relevance of oxidative stress to yeast, but it did lead to weak conversions.155 In S. enterica,
nickel toxicity in microorganisms. Many studies have shown high intracellular nickel did not result in base pair or frame-
that cells subjected to oxidative stress exhibit enhanced DNA shift mutations,156 and the metal was found to be equally toxic
damage, protein impairment, and lipid peroxidation along to wild-type E. coli cells and recA mutants.157 In contrast,
with increased titers of oxidative stress defense systems.131,132 nickel generated an increase in the mutation rate of
Below, we summarize the types of nickel-dependent oxidative Corynebacterium.158 Studies in mammals are beginning to
damage noted in vitro and describe a few examples where demonstrate that the role of nickel in carcinogenesis may be
in vivo signatures of oxidative damage by nickel are reported. due to inhibition of DNA repair and not to an acceleration of
damage.85,86,159 Similarly, nickel serves as a comutagen with
3.3.1 In vitro studies. Nickel is a rather poor generator of alkylating agents in microorganisms,160 likely due to inhibition
oxidative damage when compared to iron or copper;133 never- of DNA repair.160,161 Indeed, nickel inhibits E. coli MutT, but
theless, this metal can modify many types of biological com- with an IC50 of 1.5 mM.162 Nickel also was shown to inhibit
pounds. For example, nickel can participate in DNA various DNA polymerases in vitro with IC50’s ranging from 10
cleavage,134,135 intrastrand cross-links,136 deglycosylation,137 (without magnesium) to 780 mM.114 These data suggest that
and base modification.136–141 Furthermore, nickel can lead to nickel does not cause a significant amount of oxidative DNA
protein cleavage142 or oxidize side chains to produce Tyr-Tyr damage in vivo, and that any mutagenic effect of nickel is likely
cross-linkings.143 In addition, this metal can stimulate lipid due to inhibition of DNA repair.
oxidation in the presence of thiols144 or when combined with Oxidation of proteins has not been reported to be a major
iron.145 For such reactions, nickel uses a variety of oxidants mechanism of nickel toxicity in microorganisms. Nevertheless,
including oxygen,134,141,144 superoxide,146 hydrogen nickel increases the GSSG/GSH ratio154 which hints at the
135–138,140,146–149
peroxide, and organic hydroperoxides.139,144,150 possibility of nickel-dependent modulation of the oxidation
Of great importance when considering the chemistry of nickel, status of cysteinyl thiol in proteins. The importance of this
the metal-associated ligands play critical roles. Cellular nickel effect is unclear, however, since the levels of GSH reductase
does not exist in a fully aquated state, and the variety of were not enhanced upon nickel exposure.154
adventitious ligands which bind nickel can have a profound Oxidative stress has been shown to cause lipid peroxidation.163
impact on the ability of the metal to produce reactive oxygen Nickel-toxified filamentous fungus Curvularia lunata had an
species. Indeed, nickel bound to cysteine,144,148 histidine,148 increase in membrane lipid saturation and an enhanced level
glutathione,139,148,150 small peptides,146,147,150 and amine- of thiobarbituric acid reactive substances (TBARS).164
containing compounds144,149 accelerated the rates at which TBARS are lipid degradation products used as an indirect
nickel creates oxidative damage. On the other hand, the ability measure of lipid peroxidation. Extensive lipid peroxidation

This journal is c The Royal Society of Chemistry 2011 Metallomics, 2011, 3, 1153–1162 1159
View Article Online

only occurs in membranes containing polyunsaturated fatty unknown and the target will likely vary with the specific
acids;131,165 therefore, lipid peroxidation is unlikely to occur in physiology of the cell of interest.
most eubacteria. The requirement for polyunsaturated fatty
acids to propagate lipid peroxidation could explain the nickel- Acknowledgements
dependent increase in lipid saturation level of C. lunata.
In summary, the overall importance of oxidative stress in Studies related to this topic in the Hausinger laboratory are
nickel toxicity may be overstated for most microorganisms. supported by NIH grant DK045686.
For example, oxidative stress cannot account for nickel toxi-
city of cells grown in the absence of oxygen.166–168
Published on 28 July 2011. Downloaded by University of Florida Libraries on 27/12/2017 07:28:16.

References
1 J. O. Nriagu, Nickel in the Environment, John Wiley & Sons,
4. Perspectives New York, NY, 1980.
2 M. I. Ansari and A. Malik, Environ. Monit. Assess., 2010, 167,
Our discussion of nickel toxicity began by enumerating four 151–163.
3 T. M. Nieminen, L. Ukonmaanaho, N. Rausch and W. Shotyk, in
mechanistic hypotheses to account for microbial damage by
Metal Ions in Life Sciences, ed. A. Sigel, H. Sigel and R. K. O.
this metal. We described several examples where substitu- Sigel, John Wiley & Sons, Ltd., New York, NY, 2007, vol. 2,
tion of nickel for the normal metal leads to inactivation of a pp. 1–30.
metalloenzyme. Depending on the specific growth condi- 4 E. Denkhaus and K. Salnikow, Crit. Rev. Oncol. Hematol., 2002,
42, 35–56.
tions, such inhibition could account for some of the cellular 5 K. S. Kasprzak, F. W. Sunderman Jr. and K. Salnikow, Mutat.
toxicity by nickel. For example, nickel substitution into a Res., Fundam. Mol. Mech. Mutagen., 2003, 533, 67–97.
particular iron oxygenase could hinder growth on the 6 K. K. Das, S. N. Das and S. A. Dhundasi, Indian J. Med. Res.,
cognate carbon source. Similarly, we provided samples of 2008, 128, 412–425.
7 K. S. Kasprzak and K. Salnikow, in Metal Ions in Life Sciences,
non-metalloenzymes that are inhibited by nickel binding to ed. A. Sigel, H. Sigel and R. K. O. Sigel, John Wiley & Sons, Ltd.,
the active site. Nickel is known to bind to Cys and/or His New York, NY, 2007, vol. 2, pp. 619–660.
catalytic residues of several enzymes, with consequent 8 M. Yusuf, Q. Fariduddin, S. Hayat and A. Ahmad, Bull. Environ.
Contam. Toxicol., 2011, 86, 1–17.
reductions in activities. In addition we mentioned several 9 H. Babich and G. Stotzky, Adv. Appl. Microbiol., 1983, 29,
enzyme cases where the inhibitory metal appears to bind to 195–265.
a secondary or allosteric site to affect activity. Much of the 10 A. Rodrigue, G. Effantin and M. A. Mandrand-Berthelot,
scarce literature on the mechanisms of nickel inhibition J. Bacteriol., 2005, 187, 2912–2916.
11 D. G. Barceloux, Clin. Toxicol., 1999, 37, 239–258.
relates to the interaction of the metal with specific enzymes; 12 S. B. Mulrooney and R. P. Hausinger, FEMS Microbiol. Rev.,
however, nickel toxicosis has yet to be linked to the failure 2003, 27, 239–261.
of a specific enzyme. 13 Y. Li and D. B. Zamble, Chem. Rev., 2009, 109, 4617–4643.
14 S. W. Ragsdale, J. Biol. Chem., 2010, 284, 18571–18575.
In contrast to the many examples described above where
15 H. Kaluarachchi, K. C. C. Chung and D. B. Zamble, Nat. Prod.
nickel toxicity involves direct inhibition of enzymes, the Rep., 2010, 27, 681–694.
evidence related to the fourth hypothesis, that nickel directly 16 H. Guldan, R. Sterner and P. Babinger, Biochemistry, 2008, 47,
causes oxidative stress, is much less compelling. Future work 7376–7384.
17 H. Merkens, R. Kappl, R. P. Jakob, F. X. Schmid and S. Fetzner,
will need to determine if SOD protects E. coli from nickel in Biochemistry, 2008, 47, 12185–12196.
the absence of oxygen. For example, oxygen-independent 18 N. S. Dosanjh and S. L. J. Michel, Curr. Opin. Chem. Biol., 2006,
protection by SOD has been demonstrated for copper where 10, 123–130.
19 J. S. Iwig and P. T. Chivers, Nat. Prod. Rep., 2010, 27, 658–667.
heterologous expression of the human CuZnSOD in E. coli
20 A. Dannielli and V. Scarlato, FEMS Microbiol. Rev., 2010, 34,
protected cells from anaerobic copper overload.169 Further- 738–752.
more, MnSOD has been shown to bind metals other than 21 H. Reyes-Caballero, G. C. Campanello and D. P. Giedroc,
manganese;170,171 thus MnSOD may act as an adventitious Biophys. Chem., 2011, 156, 103–114.
22 C. Blériot, G. Effantin, F. Lagarde, M. A. Mandrand-Berthelot
nickel scavenger. If nickel toxicity does indeed lead to oxida- and A. Rodrique, J. Bacteriol., 2011, 193, 3785–3793.
tive stress, then it will be necessary to determine whether this is 23 L. A. Finney and T. V. O’Halloran, Science, 2003, 300, 931–936.
a direct or indirect effect. A likely mechanism by which nickel 24 C. E. Outten and T. V. O’Halloran, Science, 2001, 292,
could indirectly cause oxidative stress is through the increase 2488–2492.
25 K. De Pina, V. Desjardin, M. A. Mandrand-Berthelot,
of cellular iron. Iron has long been shown to be the primary G. Giordano and L. F. Wu, J. Bacteriol., 1999, 181, 670–674.
metal involved with oxidative stress. The increased demand for 26 P. T. Chivers and R. T. Sauer, Protein Sci., 1999, 8, 2494–2500.
iron import during nickel overload92 coupled to nickel’s ability 27 E. R. Schreiter, M. D. Sintchak, Y. Guo, P. T. Chivers,
R. T. Sauer and C. L. Drennan, Nat. Struct. Biol., 2003, 10,
to damage iron proteins (section 3.1.1) suggest that nickel- 794–799.
toxicity leads to an increase in the intracellular free iron pool. 28 S. C. Wang, A. V. Dias and D. B. Zamble, Dalton Trans., 2009,
This indirect enhancement of oxidative stress was also 2459–2466.
hypothesized for copper-generated oxidative stress in vivo.169 29 P. T. Chivers and R. T. Sauer, J. Biol. Chem., 2000, 275,
19735–19741.
In summary, the general importance of nickel as a universal 30 A. L. West, F. St. John, P. E. M. Lopes, A. D. MacKerell Jr.,
microbial toxicant is being realized due to the discovery of a E. Pozharski and S. L. J. Michel, J. Am. Chem. Soc., 2010, 132,
widely distributed nickel defense system (RcnA) and the 14447–14456.
prevalence of nickel in industrial pollution. Despite the 31 E. L. Benanti and P. T. Chivers, J. Biol. Chem., 2007, 282,
20365–20375.
numerous demonstrations of nickel toxicosis in all forms of 32 N. S. Dosanjh, N. A. Hammerbacher and S. L. J. Michel,
life, the mechanism of toxicity in microorganisms is generally Biochemistry, 2007, 46, 2520–2529.

1160 Metallomics, 2011, 3, 1153–1162 This journal is c The Royal Society of Chemistry 2011
View Article Online

33 C. Muller, C. Bahlawane, S. Aubert, C. M. Delay, K. Schauer, 64 T. Schmidt and H. G. Schlegel, FEMS Microbiol. Lett., 1989, 62,
I. Michaud-Soret and H. De Reuse, Nucleic Acids Res., 2011, 315–328.
DOI: 10.1093/nar/gkr460, in press. 65 L. Diels, S. Van Roy, S. Taghavi and R. Van Houdt, Antonie van
34 J. S. Iwig, J. L. Rowe and P. T. Chivers, Mol. Microbiol., 2006, Leeuwenhoek, 2009, 96, 247–258.
62, 252–262. 66 F. N. Stähler, S. Odenbreit, R. Haas, J. Wilrich, A. H. Van Vliet,
35 D. R. Campbell, K. E. Chapman, K. J. Waldron, S. Tottey, J. G. Kusters, M. Kist and S. Bereswill, Infect. Immun., 2006, 74,
S. Kendall, G. Cavallaro, C. Andreini, J. Hinds, N. G. Stoker, 3845–3852.
N. J. Robinson and J. S. Cavet, J. Biol. Chem., 2007, 282, 67 H. Liesegang, K. Lemke, R. A. Siddiqui and H. G. Schlegel,
32298–32310. J. Bacteriol., 1993, 175, 767–778.
36 J. S. Cavet, W. Meng, M. A. Pennella, R. J. Appelhoff, 68 T. Schmidt and H. G. Schlegel, J. Bacteriol., 1994, 176,
D. P. Giedroc and N. J. Robinson, J. Biol. Chem., 2002, 277, 7045–7054.
Published on 28 July 2011. Downloaded by University of Florida Libraries on 27/12/2017 07:28:16.

38441–38448. 69 C. J. Haney, G. Grass, S. Frake and C. Rensing, J. Ind. Microbiol.


37 Y. Zhang, D. A. Rodionov, M. S. Gelfand and V. N. Gladyshev, Biotechnol., 2005, 32, 215–226.
BMC Genomics, 2009, 10, 78. 70 G. Grass, B. Fan, B. P. Rosen, K. Lemke, H. G. Schlegel and
38 T. Eitinger and M. A. Mandrand-Berthelot, Arch. Microbiol., C. Rensing, J. Bacteriol., 2001, 183, 2803–2807.
2000, 173, 1–9. 71 J. E. Park, H. G. Schlegel, H. G. Rhie and H. S. Lee, Int.
39 T. Eitinger, J. Suhr, L. Moore and J. A. C. Smith, BioMetals, Microbiol., 2004, 7, 27–34.
2005, 18, 399–405. 72 D. Munkelt, G. Grass and D. H. Nies, J. Bacteriol., 2004, 186,
40 C. Navarro, L. F. Wu and M. A. Mandrand-Berthelot, Mol. 8036–8043.
Microbiol., 1993, 9, 1181–1191. 73 W. W. Tso and J. Adler, J. Bacteriol., 1974, 118, 560–576.
41 M. V. Cherrier, C. Cavazza, C. Bochot, D. Lemaire and 74 D. Borrok, M. J. Borrok, J. B. Fein and L. L. Kiessling, Environ.
J. C. Fontecilla-Camps, Biochemistry, 2008, 47, 9937–9943. Sci. Technol., 2005, 39, 5227–5233.
42 V. Jubier-Maurin, A. Rodrigue, S. Ouahrani-Bettache, 75 D. L. Englert, C. A. Adase, A. Jayaraman and M. D. Manson,
M. Layssac, M. A. Mandrand-Berthelot, S. Köhler and J. Bacteriol., 2010, 192, 2633–2637.
J. P. Liautard, J. Bacteriol., 2001, 183, 426–434. 76 M. Joho, M. Inouhe, H. Tohoyama and T. Murayama, FEMS
43 K. S. Park, T. Iida, Y. Yamaichi, T. Oyagi, K. Yamamoto and Microbiol. Lett., 1990, 66, 333–338.
T. Honda, Infect. Immun., 2000, 68, 5742–5748. 77 M. Joho, Y. Ishikawa, M. Kunikane, M. Inouhe, H. Tohoyama
44 C. S. Beckwith, D. J. McGee, H. L. T. Mobley and L. K. Riley, and T. Murayama, Microbios., 1992, 71, 149–159.
Infect. Immun., 2001, 69, 5914–5920. 78 D. A. Pearce and F. Sherman, J. Bacteriol., 1999, 181, 4774–4779.
45 F. Sebbane, M. A. Mandrand-Berthelot and M. Simonet, 79 R. Ruotolo, G. Marchini and S. Ottonello, GenomeBiology, 2008,
J. Bacteriol., 2002, 184, 5706–5713. 9, R67.
46 A. Hiron, B. Posteraro, M. Carrière, L. Remy, C. Delporte, 80 A. Arita, X. Zhou, T. P. Ellen, X. Liu, J. Bai, J. P. Rooney,
M. La Sorda, M. Sanguinetti, V. Juillard and E. Borezée-Durant, A. Kurtz, C. B. Klein, W. Dai, T. J. Begley and M. Costa, BMC
Mol. Microbiol., 2010, 77, 1246–1260. Genomics, 2009, 10, 524.
47 L. F. Wu, M. A. Mandrand-Berthelot, R. Waugh, C. J. Edmonds, 81 D. Fortin, G. Southam and T. J. Beveridge, FEMS Microbiol.
S. E. Holt and D. H. Boxer, Mol. Microbiol., 1989, 3, 1709–1718. Ecol., 1994, 14, 121–132.
48 G. Eberz, T. Eitinger and B. Friedrich, J. Bacteriol., 1989, 171, 82 V. N. Tripathi and S. Srivastava, Can. J. Microbiol., 2006, 52,
1340–1345. 287–292.
49 T. Eitinger and B. Friedrich, J. Biol. Chem., 1991, 266, 83 P. Sar, S. K. Kazy and S. P. Singh, Lett. Appl. Microbiol., 2001,
3222–3227. 32, 257–261.
50 L. Wolfram, B. Friedrich and T. Eitinger, J. Bacteriol., 1995, 177, 84 O. A. Zadvornyy, M. Allen, S. K. Brumfield, Z. Varpness,
1840–1843. E. S. Boyd, N. A. Zorin, L. Serebriakova, T. Douglas and
51 T. Eitinger and B. Friedrich, Mol. Microbiol., 1994, 12, J. W. Peters, Environ. Sci. Technol., 2010, 44, 834–840.
1025–1032. 85 H. Chen and M. Costa, BioMetals, 2009, 22, 191–196.
52 H. Komeda, M. Kobayashi and S. Shimizu, Proc. Natl. Acad. Sci. 86 H. Chen, N. C. Giri, R. Zhang, K. Yamane, Y. Zhang,
U. S. A., 1997, 94, 36–41. M. Maroney and M. Costa, J. Biol. Chem., 2010, 285, 7374–7383.
53 O. Degen, M. Kobayashi, S. Shimizu and T. Eitinger, Arch. 87 R. Sekirnik, N. R. Rose, J. Mecinović and C. J. Schofield,
Microbiol., 1999, 171, 139–145. Metallomics, 2010, 2, 397–399.
54 C. Fu, S. Javedan, F. Moshiri and R. J. Maier, Proc. Natl. Acad. 88 Z. Cheng, Y. Y. C. Wei, W. W. L. Sung, B. R. Glick and
Sci. U. S. A., 1994, 91, 5099–5103. B. J. McConkey, Proteome Sci., 2009, 7, 18.
55 H. L. T. Mobley, R. M. Garner and P. Bauerfeind, Mol. Microbiol., 89 J. Kuchar and R. P. Hausinger, Chem. Rev., 2004, 104, 509–526.
1995, 16, 97–109. 90 D. Koch, D. H. Nies and G. Grass, BioMetals, 2007, 20, 759–771.
56 S. T. Cole, R. Brosch, J. Parkhill, T. Garnier, C. Churcher, 91 J. P. McHugh, F. Rodriguez-Quiñones, H. Abdul-Tehrani,
D. Harris, S. V. Gordon, K. Eiglmeier, S. Gas, C. E. Barry, D. A. Svistunenko, R. K. Poole, C. E. Cooper and
3rd, F. Tekaia, K. Badcock, D. Basham, D. Brown, S. C. Andrews, J. Biol. Chem., 2003, 278, 29478–29486.
T. Chillingworth, R. Connor, R. Davies, K. Devlin, T. Feltwell, 92 S. Wang, Y. Wu and F. W. Outten, J. Bacteriol., 2011, 193,
S. Gentles, N. Hamlin, S. Holroyd, T. Hornsby, K. Jagels, 563–574.
A. Krogh, J. McLean, S. Moule, L. Murphy, K. Oliver, 93 R. P. Hausinger, Crit. Rev. Biochem. Mol. Biol., 2004, 39, 21–68.
J. Osborne, M. A. Quail, M. A. Rajandream, J. Rogers, 94 V. Purpero and G. R. Moran, J. Biol. Inorg. Chem., 2007, 12,
S. Rutter, K. Seeger, J. Skelton, R. Squares, S. Squares, 587–601.
J. E. Sulston, K. Taylor, S. Whitehead and B. G. Barrell, Nature, 95 N. R. Rose, M. A. McDonough, O. N. F. King, A. Kawamura
1998, 393, 537–544. and C. J. Schofield, Chem. Soc. Rev., 2011, 40, 4364–4397.
57 T. Eitinger, O. Degen, U. Böhnke and M. Müller, J. Biol. Chem., 96 E. Kalliri, P. K. Grzyska and R. P. Hausinger, Biochem. Biophys.
2000, 275, 18029–18033. Res. Commun., 2005, 338, 191–197.
58 D. Niegowski and S. Eshaghi, Cell. Mol. Life Sci., 2007, 64, 97 L. Bankel, G. Lindstedt and S. Lindstedt, J. Biol. Chem., 1972,
2564–2574. 247, 6128–6134.
59 K. Schauer, B. Gouget, M. Carrière, A. Labigne and H. de Reuse, 98 M. Costa, T. L. Davidson, H. Chen, Q. Ke, P. Zhang, Y. Yan,
Mol. Microbiol., 2007, 63, 1054–1068. C. Huang and T. Kluz, Mutat. Res., Fundam. Mol. Mech.
60 G. S. Davis, E. L. Flannery and H. L. T. Mobley, Infect. Immun., Mutagen., 2005, 592, 79–88.
2006, 74, 6811–6820. 99 Y. P. Chen, M. J. Dilworth and A. R. Glenn, Arch. Microbiol.,
61 D. A. Rodionov, P. Hebbeln, M. S. Gelfand and T. Eitinger, 1984, 138, 187–190.
J. Bacteriol., 2006, 188, 317–327. 100 Y. P. Chen, A. R. Glenn and M. J. Dilworth, Arch. Microbiol.,
62 K. Schauer, D. A. Rodionov and H. de Reuse, Trends Biochem. 1985, 141, 225–228.
Sci., 2008, 33, 330–338. 101 D. E. Sheppard, J. T. Penrod, T. Bobik, E. Kofoid and
63 D. H. Nies, FEMS Microbiol. Rev., 2003, 27, 313–339. J. R. Roth, J. Bacteriol., 2004, 186, 7635–7644.

This journal is c The Royal Society of Chemistry 2011 Metallomics, 2011, 3, 1153–1162 1161
View Article Online

102 N. R. Buan and J. C. Escalante-Semerena, J. Biol. Chem., 2006, 137 K. S. Kasprzak and L. Hernandez, Cancer Res., 1989, 49,
281, 16971–16977. 5964–5968.
103 N. R. Buan, S. J. Suh and J. C. Escalante-Semerena, J. Bacteriol., 138 P. Kaczmarek, W. Szczepanik and M. Jeźowska-Bojczuk, Dalton
2004, 186, 5708–5714. Trans., 2005, 3653–3657.
104 E. Kofoid, C. Rappleye, I. Stojiljkovic and J. Roth, J. Bacteriol., 139 X. Shi, Y. Mao, N. Ahmed and H. Jiang, J. Inorg. Biochem.,
1999, 181, 5317–5329. 1995, 57, 91–102.
105 J. L. Seffernick, H. McTavish, J. P. Osborne, M. L. de Souza, 140 M. C. Kelly, G. Whitaker, B. White and M. R. Smyth, Free
M. J. Sadowsky and L. P. Wackett, Biochemistry, 2002, 41, Radical Biol. Med., 2007, 42, 1680–1689.
14430–14437. 141 J. J. Butzow and G. L. Eichhorn, Biopolymers, 1965, 3, 95–107.
106 G. Niemirowicz, F. Parussini, F. Agüero and J. J. Cazzulo, 142 W. Bal, R. Liang, J. Lukszo, S. H. Lee, M. Dizdaroqlu and
Biochem. J., 2007, 401, 399–410. K. S. Kasprzak, Chem. Res. Toxicol., 2000, 13, 616–624.
Published on 28 July 2011. Downloaded by University of Florida Libraries on 27/12/2017 07:28:16.

107 J. Arima, M. Iwabuchi and T. Hatanaka, Biochem. Biophys. Res. 143 G. Gill, A. A. Richter-Rusli, M. Ghosh, C. J. Burrows and
Commun., 2004, 317, 531–538. S. E. Rokita, Chem. Res. Toxicol., 1997, 10, 302–309.
108 W. T. Lowther and B. W. Matthews, Chem. Rev., 2002, 102, 144 X. Shi, N. S. Dalal and K. S. Kasprzak, J. Inorg. Biochem., 1993,
4581–4608. 15, 211–225.
109 B. Chevrier, H. D’Orchymont, C. Schalk, C. Tarnus and 145 M. G. Repetto, N. F. Ferrarotti and A. Boveris, Arch. Toxicol.,
D. Moras, Eur. J. Biochem., 1996, 237, 393–398. 2010, 84, 255–262.
110 M. Simitsopoulou, A. Vafopoulou, T. Choli-Papadopoulou and 146 N. Cotelle, E. Trémolières, J. L. Bernier, J. P. Catteau and
E. Alichanidis, Appl. Environ. Microbiol., 1997, 63, 4872–4876. J. P. Hénichart, J. Inorg. Biochem., 1992, 46, 7–15.
111 J. M. Dolence, L. E. Steward, E. K. Dolence, D. H. Wong and 147 J. Torreilles and M. C. Guérin, FEBS Lett., 1990, 272, 58–60.
C. D. Poulter, Biochemistry, 2000, 39, 4096–4104. 148 S. Joshi, M. M. Husain, R. Chandra, S. K. Hasan and
112 K. Sato, A. Okubo and S. Yamazaki, J. Biochem., 1998, 124, 51–54. R. C. Srivastava, Hum. Exp. Toxicol., 2005, 24, 13–17.
113 L. Debussche, M. Couder, D. Thibaut, B. Cameron, J. Crouzet 149 H. Sigel, K. Wyss, P. Waldmeier and R. Griesser, J. Coord.
and F. Blanche, J. Bacteriol., 1992, 174, 7445–7451. Chem., 1974, 3, 235–247.
114 E. T. Snow, L. S. Xu and P. L. Kinney, Chem.-Biol. Interact., 150 X. Shi, N. S. Dalal and K. S. Kasprzak, Arch. Biochem. Biophys.,
1993, 88, 155–173. 1992, 299, 154–162.
115 T. Nakai, T. Hasegawa, E. Yamashita, M. Yamamoto, 151 J. D. Van Nostrand, J. M. Arthur, L. E. Kilpatrick, B. A. Neely,
T. Kumasaka, T. Ueki, H. Nanba, Y. Ikenaka, S. Takahashi, P. M. Bertsch and P. J. Morris, Microbiology, 2008, 154,
M. Sato and T. Tsukihara, Structure, 2000, 8, 729–737. 3813–3824.
116 W. C. Wang, W. H. Hsu, F. T. Chien and C. Y. Chen, J. Mol. 152 C. Geslin, J. Llanos, D. Prieur and C. Jeanthon, Res. Microbiol.,
Biol., 2001, 306, 251–261. 2001, 152, 901–905.
117 A. Louwrier and C. J. Knowles, Enzyme Microb. Technol., 1996, 153 W. B. Healy, S. C. Cheng and W. D. McElvoy, Arch. Biochem.
19, 562–571. Biophys., 1955, 54, 206–214.
118 A. J. Barrett and N. D. Rawlings, Biol. Chem., 2001, 382, 154 V. K. Randhawa, F. Zhou, X. Jin, C. Nalewajko and
727–733. D. J. Kushner, Can. J. Microbiol., 2001, 47, 987–993.
119 P. M. Cummins and B. O’Connor, Int. J. Biochem. Cell Biol., 155 I. Singh, Mutat. Res., Genet. Toxicol., 1984, 137, 47–49.
1996, 28, 883–893. 156 N. W. Biggart and M. Costa, Mutat. Res. Lett., 1986, 175,
120 M. Schaeffer, A. de Miranda, J. C. Mottram and G. H. Coombs, 209–215.
Mol. Biochem. Parasitol., 2006, 150, 318–329. 157 G. Brandi, G. F. Schiavano, A. Albano, F. Cattabeni and
121 T. Barkay, S. M. Miller and A. O. Summers, FEMS Microbiol. O. Cantoni, Mutat. Res. Lett., 1990, 245, 201–204.
Rev., 2003, 27, 355–384. 158 P. Pikálek and J. Necásek, Folia Microbiol., 1983, 28, 17–21.
122 S. Ghosh, P. C. Sadhukhan, J. Chaudhuri, D. K. Ghosh and 159 A. Hartwig and D. Beyersmann, Mutat. Res., 1989, 217, 65–73.
A. Mandal, J. Appl. Microbiol., 1999, 86, 7–12. 160 J. S. Dubins and J. M. LaVelle, Mutat. Res., Fundam. Mol. Mech.
123 J. Hurwitz, M. Gold and M. Anders, J. Biol. Chem., 1964, 239, Mutagen., 1986, 162, 187–199.
3474–3482. 161 K. Woźniak and J. B"asiak, Cell Mol. Biol. Lett., 2004, 9,
124 J. Urbonavičius, G. Jäger and G. R. Björk, Nucleic Acids Res., 83–94.
2007, 35, 3297–3305. 162 D. W. Porter, H. Yakushiji, Y. Nakabeppu, M. Sekiguchi,
125 Y. Liu and D. V. Santi, Proc. Natl. Acad. Sci. U. S. A., 2000, 97, M. J. Fivash Jr. and K. S. Kasprzak, Carcinogenesis, 1997, 18,
8263–8265. 1785–1791.
126 S. Sommaruga, L. De Gioia, P. Tortora and A. Polissi, Biochem. 163 E. Niki, Y. Yoshida, Y. Saito and N. Noguchi, Biochem. Biophys.
Biophys. Res. Commun., 2009, 388, 222–227. Res. Commun., 2005, 338, 668–676.
127 T. C. Meredith and R. W. Woodard, J. Biol. Chem., 2003, 278, 164 K. Paraszkiewicz, P. Bernat, M. Na"iwajski and J. D"ugoński,
32771–32777. Arch. Microbiol., 2010, 192, 135–141.
128 G. P. A. Bongaerts, J. Uitzetter, R. Brouns and G. D. Vogels, 165 B. H. Bielski, R. L. Arudi and M. W. Sutherland, J. Biol. Chem.,
Biochim. Biophys. Acta, 1978, 527, 348–358. 1983, 258, 4759–4761.
129 R. D. Imhoff, N. P. Power, M. J. Borrok and P. A. Tipton, 166 L. F. Wu, C. Navarro, K. de Pina, M. Quénard and M. A.
Biochemistry, 2003, 42, 4094–4100. Mandrand, Environ. Health Perspect., 1994, 102(Suppl 3),
130 L. Gabison, M. Chiadmi, M. El Hajji, B. Castro, N. Colloc’h and 297–300.
T. Prangé, Acta Crystallogr., Sect. D: Biol. Crystallogr., 2010, 66, 167 G. Zayed and J. Winter, Appl. Microbiol. Biotechnol., 2000, 53,
714–724. 726–731.
131 J. A. Imlay, Annu. Rev. Microbiol., 2003, 57, 395–418. 168 J. Bartacek, F. G. Fermoso, A. B. Catena and P. N. L. Lens,
132 J. A. Imlay, Annu. Rev. Biochem., 2008, 77, 755–776. J. Hazard. Mater., 2010, 180, 289–296.
133 M. Valko, H. Morris and M. T. Cronin, Curr. Med. Chem., 2005, 169 L. Macomber and J. A. Imlay, Proc. Natl. Acad. Sci. U. S. A.,
12, 1161–1208. 2009, 106, 8344–8349.
134 C.-C. Cheng, S. E. Rokita and C. J. Burrows, Angew. Chem., Int. 170 W. F. Beyer Jr. and I. Fridovich, J. Biol. Chem., 1991, 266,
Ed. Engl., 1993, 32, 277–278. 303–308.
135 S. Kawanishi, S. Inoue and K. Yamamoto, Carcinogenesis, 1989, 171 M. M. Whittaker and J. W. Whittaker, Biochemistry, 1997, 36,
10, 2231–2235. 8923–8931.
136 D. R. Lloyd and D. H. Phillips, Mutat. Res., Fundam. Mol. Mech. 172 J. R. O’Brien, D. J. Schuller, V. S. Yang, B. D. Dillard and
Mutagen., 1999, 424, 23–36. W. N. Lanzilotta, Biochemistry, 2003, 42, 5547–5554.

1162 Metallomics, 2011, 3, 1153–1162 This journal is c The Royal Society of Chemistry 2011

You might also like