TY - JOUR AU1 - Barkay,, Tamar AU2 - Miller, Susan, M. AU3 - Summers, Anne, O. AB - Abstract Bacterial resistance to inorganic and organic mercury compounds (HgR) is one of the most widely observed phenotypes in eubacteria. Loci conferring HgR in Gram-positive or Gram-negative bacteria typically have at minimum a mercuric reductase enzyme (MerA) that reduces reactive ionic Hg(II) to volatile, relatively inert, monoatomic Hg(0) vapor and a membrane-bound protein (MerT) for uptake of Hg(II) arranged in an operon under control of MerR, a novel metal-responsive regulator. Many HgR loci encode an additional enzyme, MerB, that degrades organomercurials by protonolysis, and one or more additional proteins apparently involved in transport. Genes conferring HgR occur on chromosomes, plasmids, and transposons and their operon arrangements can be quite diverse, frequently involving duplications of the above noted structural genes, several of which are modular themselves. How this very mobile and plastic suite of proteins protects host cells from this pervasive toxic metal, what roles it has in the biogeochemical cycling of Hg, and how it has been employed in ameliorating environmental contamination are the subjects of this review. Metalloregulation, Flavin disulfide oxidoreductase, Metallochaperone, Metal transport, Global mercury cycle, Horizontal gene exchange, Bioremediation 1 Introduction Study of bacterial mercury resistance is in its fifth decade. This earliest discovered example of a bacterial defense system against this ubiquitous toxic metal and its organic derivatives has been studied for its novel biochemical and genetic aspects, for its natural and engineered role in environmental bioremediation and biomonitoring, as a model of horizontal gene exchange, and for the general insights it provides into how cells deal with redox-active thiophilic metals. The genes encoding the proteins of mercury resistance occur naturally on chromosomes, plasmids, and transposable elements in a striking diversity of arrangements, often involving duplications and distributions of the enzymes, transporters or regulators among several replicons in one cell. Moreover, two major mer genes, the regulator MerR and the major detoxification enzyme, MerA, the mercuric ion reductase, are each composed of discrete modules observed in paralogs with distinct but related roles in prokaryotic physiology. This brief review touches on each of these aspects of this appropriately protean system which has evolved in response to an equally protean and unremittingly toxic element. We endeavor to place the coordinated workings of the mer proteins into a cellular context; a theme throughout will be the roles of cysteines in the mer proteins and their relationship to thiols in their respective cellular compartments. Cysteines are the most vulnerable targets of mercury-induced loss of function in any protein. Surprisingly, rather than dodging such interactions, every mer protein uses cysteines (perhaps exclusively) to interact with either inorganic or organic mercurials. Thus, mer proteins have evolved to carry out metalloregulation, transport, and enzyme catalysis but emerge unscathed from their interactions with mercury, converting the tenaciously reactive ionic or organic forms into the reduced and relatively inert, monoatomic elemental form, Hg(0). Since Hg(0) has a high vapor pressure (Henry's constant of 0.3) and very low aqueous solubility (6 μg per 100 ml of water at 25°C), it is volatile at a liquid:air interface but may coalesce into a liquid in a closed or diffusion-limited system. We are now beginning to understand how the mer proteins have evolved to exploit the internal cellular milieu to manage these feats. Mercury resistance is also the only bacterial metal resistance system whose mechanism leads to large-scale transformation of its toxic target. The mechanisms of other cation and oxoanion resistances are based on efflux pumps or extracellular sequestration. Thus, there has long been interest in mer's role in the global cycling of mercury and in employing the resistance mechanism for remediation efforts. Most of these practical applications are as yet only demonstration projects, but a few have become economically viable. 1.1 A brief history of the study of mercury resistance Moore [1] first reported bacterial resistance to inorganic and organic mercury compounds (HgR) in a clinical isolate of Staphylococcus aureus which was also penicillin-resistant. This was of concern because at the time mercurial compounds were extensively used as topical disinfectants and antiseptics in hospitals and in the community. Shortly thereafter groups in the UK [2] and the USA [3] demonstrated that both HgR and penicillin resistance in S. aureus strains were genetically linked on a newly discovered class of mobile genetic elements called plasmids. Similar findings were made in clinical strains of Escherichia coli[4]. About the same time owing to growing concern about the use of mercurial fungicides on grain in Sweden and elsewhere [5, 6], there were several reports of non-clinical, environmental bacteria which could transform organomercurials into volatile, relatively inert monoatomic Hg(0) [7–9]. It was quickly recognized that some bacteria were resistant to and able to transform both inorganic and organic mercury compounds and other bacteria (even of the same genus) could only resist and transform inorganic mercury compounds [10]. The former phenotype was given the name ‘broad-spectrum Hg resistance’ and the latter ‘narrow-spectrum Hg resistance’ [11]. Work on the mechanism of HgR led to identification of the key detoxification enzyme, MerA, the mercuric ion reductase [12], and also of a second enzyme, MerB, which split the carbon–Hg bond in such compounds as the disinfectant phenylmercuric acetate (PMA) and the fungicide methylmercury chloride, a potent neurotoxic agent [13]. Membrane and periplasmic proteins involved in the seemingly paradoxical inward transport of ionic mercury [14, 15] were also identified [16, 17]. The first sequences of HgR loci revealed proteins corresponding to these biochemical and physiological functions as well as a candidate regulatory gene (merR). Somewhat surprisingly, each operon had distinct gene complements and operonic arrangements [18–21]. Later sequencing efforts showed even greater variety in contents and arrangements of mer operons [22–27]. At present, as detailed below, considerable recent progress has been made in understanding the structure–function relationships in many of these genes. The population biology of HgR is more extensive than that for other bacterial metal resistances owing to concerns over the environmental dissemination of Hg. Aerobic and facultative HgR bacteria are easily isolated on a variety of media from soils, water, and sediments as well as from humans and other animals [23, 25, 28–36]. In cases where it has been examined these bacteria reduce Hg(II) to Hg(0). In general, but not always, DNA similar to either widely found Gram-negative or Gram-positive mer genes (typically MerA) can be found in the genomes or plasmids of these strains. Thus, it appears that while Hg(II) reduction is a common mechanism, there may be even greater divergence in the mechanisms of this reduction than is so far apparent. Details and general themes arising from this great phylogenetic variety are described below. 2 The global mercury cycle 2.1 Sources of Hg in the environment Mercury is present in the Earth's crust at concentrations ranging from 21 (lower crust) to 56 (upper crust) ppb [37] in the elemental form and as a variety of HgS binary minerals, such as cinnabar, metacinnabar and hypercinnabar [38]. Although these minerals have a low solubility (10−6 g per 100 ml water for cinnabar [39]), their solubilized mercury products are available for transformations between the +2, +1 and 0 oxidation states and for conversion to various organic forms by natural and anthropogenic processes. Anthropogenic sources of Hg include input to the atmosphere resulting from burning fossil fuels and incineration or other disposal of products such as fluorescent light fixtures, batteries, dental restorations, and electrodes used in the chlor-alkali process. Deposition of Hg to land and natural waters can result from existing and defunct industries, from landfills and as a result of sludge applications [40]. Worldwide anthropogenic input is substantial, accounting for 75% of the global input of Hg to the environment [41]. Changes in atmospheric Hg concentrations over time [42, 43] and enhanced accumulation rates inferred from sediment [44, 45] and ice cores [46] clearly show increasing Hg input to the environment in the last two centuries. Biotic and abiotic processes facilitate Hg cycling from soils and water to the atmosphere and back to the surface by both wet and dry deposition (Fig. 1 and Table 1) [47]. The following is a brief discussion of the contribution of these biotic and abiotic transformations that constitute the global Hg cycle. Figure 1 Open in new tabDownload slide The biogeochemical cycle of mercury in the environment (modified from [47]). Solid arrows represent transformation or uptake of mercury. Hollow arrows indicate flux of mercury between different compartments in the environment. The width of the hollow arrows is approximately proportional to the relative importance of the flux in nature. The speciation of Hg(II) in oxic and anoxic waters is controlled by chloride and hydroxide, and by sulfide respectively [47]. Transformations known to be mediated by microorganisms are represented by circles depicting bacterial cells. SRB stands for sulfate-reducing bacteria, and merB and merA refer to the activity of genes encoding the enzymes organomercurial lyase and mercuric reductase, respectively. A group of dots indicate the involvement of unicellular algae. Light-mediated water column transformations are positioned below the sun. Photodegradation of CH3Hg+ results in mostly Hg° and an unknown C1 species depicted as Cx (D. Krabbenhoft, personal communication). Reprinted with permission from [77]. Figure 1 Open in new tabDownload slide The biogeochemical cycle of mercury in the environment (modified from [47]). Solid arrows represent transformation or uptake of mercury. Hollow arrows indicate flux of mercury between different compartments in the environment. The width of the hollow arrows is approximately proportional to the relative importance of the flux in nature. The speciation of Hg(II) in oxic and anoxic waters is controlled by chloride and hydroxide, and by sulfide respectively [47]. Transformations known to be mediated by microorganisms are represented by circles depicting bacterial cells. SRB stands for sulfate-reducing bacteria, and merB and merA refer to the activity of genes encoding the enzymes organomercurial lyase and mercuric reductase, respectively. A group of dots indicate the involvement of unicellular algae. Light-mediated water column transformations are positioned below the sun. Photodegradation of CH3Hg+ results in mostly Hg° and an unknown C1 species depicted as Cx (D. Krabbenhoft, personal communication). Reprinted with permission from [77]. Table 1 Transformations in the Hg biogeochemical cycle Transformation Process Mechanism References Hg(II) methylation Biotic Methyltransferase-mediated transfer of CH3− from methylcobalamin via the acetyl-CoA pathway in sulfate-reducing bacteria [63] Abiotic Methylation by fulvic and humic acids, carboxylic acids and methylated tin compounds [72, 73] CH3Hg(I) demethylation Biotic Reductive demethylation by merAB yielding CH4 and Hg(0). Oxidative demethylation by (an) uncharacterized microbial process(es) yielding CO2 and an unidentified Hg moiety [75, 238, 239] Abiotic Photodegradation at 200–400 nm [81, 82] Hg(II) reduction Biotic Bacterial mercuric reductase and undefined pathways by algae during light and dark growth [95, 100] Abiotic Photochemical and dark reactions by organic and inorganic free radicals. Disproportionation: 2Hg(I)→Hg(II)+Hg(0) [83, 104] Hg(0) oxidation Biotic Oxidation by hydroperoxidases in microbes, plants and animals [105, 109] Abiotic Photooxidation by various oxidants and free radicals and dark oxidation possibly by O2 in saline conditions [110, 113] Transformation Process Mechanism References Hg(II) methylation Biotic Methyltransferase-mediated transfer of CH3− from methylcobalamin via the acetyl-CoA pathway in sulfate-reducing bacteria [63] Abiotic Methylation by fulvic and humic acids, carboxylic acids and methylated tin compounds [72, 73] CH3Hg(I) demethylation Biotic Reductive demethylation by merAB yielding CH4 and Hg(0). Oxidative demethylation by (an) uncharacterized microbial process(es) yielding CO2 and an unidentified Hg moiety [75, 238, 239] Abiotic Photodegradation at 200–400 nm [81, 82] Hg(II) reduction Biotic Bacterial mercuric reductase and undefined pathways by algae during light and dark growth [95, 100] Abiotic Photochemical and dark reactions by organic and inorganic free radicals. Disproportionation: 2Hg(I)→Hg(II)+Hg(0) [83, 104] Hg(0) oxidation Biotic Oxidation by hydroperoxidases in microbes, plants and animals [105, 109] Abiotic Photooxidation by various oxidants and free radicals and dark oxidation possibly by O2 in saline conditions [110, 113] Open in new tab Table 1 Transformations in the Hg biogeochemical cycle Transformation Process Mechanism References Hg(II) methylation Biotic Methyltransferase-mediated transfer of CH3− from methylcobalamin via the acetyl-CoA pathway in sulfate-reducing bacteria [63] Abiotic Methylation by fulvic and humic acids, carboxylic acids and methylated tin compounds [72, 73] CH3Hg(I) demethylation Biotic Reductive demethylation by merAB yielding CH4 and Hg(0). Oxidative demethylation by (an) uncharacterized microbial process(es) yielding CO2 and an unidentified Hg moiety [75, 238, 239] Abiotic Photodegradation at 200–400 nm [81, 82] Hg(II) reduction Biotic Bacterial mercuric reductase and undefined pathways by algae during light and dark growth [95, 100] Abiotic Photochemical and dark reactions by organic and inorganic free radicals. Disproportionation: 2Hg(I)→Hg(II)+Hg(0) [83, 104] Hg(0) oxidation Biotic Oxidation by hydroperoxidases in microbes, plants and animals [105, 109] Abiotic Photooxidation by various oxidants and free radicals and dark oxidation possibly by O2 in saline conditions [110, 113] Transformation Process Mechanism References Hg(II) methylation Biotic Methyltransferase-mediated transfer of CH3− from methylcobalamin via the acetyl-CoA pathway in sulfate-reducing bacteria [63] Abiotic Methylation by fulvic and humic acids, carboxylic acids and methylated tin compounds [72, 73] CH3Hg(I) demethylation Biotic Reductive demethylation by merAB yielding CH4 and Hg(0). Oxidative demethylation by (an) uncharacterized microbial process(es) yielding CO2 and an unidentified Hg moiety [75, 238, 239] Abiotic Photodegradation at 200–400 nm [81, 82] Hg(II) reduction Biotic Bacterial mercuric reductase and undefined pathways by algae during light and dark growth [95, 100] Abiotic Photochemical and dark reactions by organic and inorganic free radicals. Disproportionation: 2Hg(I)→Hg(II)+Hg(0) [83, 104] Hg(0) oxidation Biotic Oxidation by hydroperoxidases in microbes, plants and animals [105, 109] Abiotic Photooxidation by various oxidants and free radicals and dark oxidation possibly by O2 in saline conditions [110, 113] Open in new tab 2.2 Ionic mercury [Hg(II)] methylation Owing to some devastating historical episodes of mass intoxication resulting from very high exposures to anthropogenic methylmercury (MeHg), current public health concerns focus primarily on this potent neurotoxin [40, 48, 49]. Although the bulk of Hg in the environment is in inorganic forms [50], natural biotic and abiotic processes can synthesize MeHg that is then subject to bioaccumulation unless degraded by other biotic or abiotic processes [51]. Discovery of Hg methylation by microorganisms in anaerobic sediments [52] focused attention on methylcorrinoids (vitamin B12) as the most likely agent able to transfer CH3− to Hg2+[53]. Research on corrinoid-producing anaerobes including methanogens [54] was followed by the discovery that many bacteria can methylate Hg(II) in lab culture [55]. Subsequent field and laboratory experiments with anaerobic sediments using specific metabolic inhibitors and substrates clearly implicated sulfate-reducing bacteria (SRB) as the principal methylators in natural anaerobic sediments [56–58]. More recent work, combining Hg chemical speciation with methylation rate measurements [59–61], showed that soluble, neutral HgS is the substrate for methylation by SRB. Richard Bartha and his students, studying methylation by crude cell extracts of Desulfovibrio desulfuricans strain LS, showed that the methyl group originated either from serine C3 [62] or from formate via acetyl-CoA [63] via CH3-tetrahydrofolate (MeTHF) to methylcobalamin [64] followed by enzymatic methylation of Hg [65]. A positive correlation of Hg(II)-dependent transformation of MeTHF to THF in soil and sediment extracts with in situ MeHg concentrations [66] further supported the notion that enzymatic catalysis, rather than spontaneous transfer of CH3 from methylcobalamin [53], is the mechanism of microbial MeHg synthesis. The microbial production of dimethylmercury (diMeHg), although proposed initially [54], has not been unequivocally demonstrated to date. The production of diMeHg from monoMeHg [67] by disproportionation in H2S-rich environments [68] might be the mechanism by which SRB form diMeHg during sulfidogenic growth [69]. Recent reports of high levels of diMeHg in aerial fluxes from oceanic upwelling sites [70] (N. Bloom, personal communication) and terrestrial [71] sources warrant a closer look at the mechanisms by which diMeHg is formed. Abiotic methylation of Hg occurs and its contribution to MeHg production in the environment is presently a contested issue [72]. Agents shown to be responsible for abiotic Hg methylation include humic and fulvic acids [72], carboxylic acids [73], and alkylated tin compounds used in agriculture as fungicides and as marine antifouling agents [74]. 2.3 CH3Hg(I) demethylation MeHg can be degraded reductively to CH4 and elemental Hg, Hg(0), by mer operon functions (see below). Alternatively, under certain conditions oxidative demethylation, the degradation of MeHg to CO2 and a small amount of CH4, occurs, possibly as a cometabolic by-product of methylotrophic metabolism [75]. Like methylation, oxidative demethylation is mediated by anaerobic bacteria. Its mechanism is currently unknown, but may be analogous to monomethylamine degradation by methanogens or to acetate oxidation by SRB [76]. Whether MeHg is degraded reductively or oxidatively in anaerobic sediments may be important. The reductive, mer operon-determined process results in net removal of Hg from the sediment as monoatomic elemental vapor. In contrast, Hg(II), the probable product of oxidative demethylation, is a substrate for re-methylation within the sediment community. Thus, a methylation–demethylation cycle may exist in environments lacking the mer-mediated process [77]. Several environmental studies [75, 78, 79] suggest that reductive mer-mediated demethylation dominates at high Hg concentrations in more aerobic settings, whereas oxidative demethylation dominates at lower Hg concentrations in more anaerobic settings (see [77] for details). That this pattern might arise from the effect of oxygen on the induction of mer operon expression is suggested by the observation that expression of the broad-spectrum mer operon of a denitrifier, Pseudomonas stutzeri OX, [80], was induced at lower Hg(II) concentrations during aerobic growth than during anaerobic growth [77]. Thus, conditional inducibility of the mer operon may critically affect MeHg production in Hg-contaminated environments. Abiotic degradation of MeHg can be effected by sunlight, specifically UV-A and UV-B, spanning a wavelength range of 280–400 nm, and this process is inhibited by singlet oxygen-trapping agents [81]. Dark incubations of lake water had 350-fold higher MeHg concentrations than light incubations [82]. Suda et al. [81] reported production of inorganic Hg from MeHg during photodegradation and Hg(0) was identified as the major product of photodegradation in wetlands (D. Krabbenhoft, personal communication). Thus, in light-exposed environments, such as wetlands and lakes, and especially at low total Hg concentrations, photodegradation may be the major mechanism for MeHg degradation [82]. In contrast, in sediments and bottom waters, where MeHg accumulates following methylation, photodegradation may have little impact on demethylation and either reductive or oxidative microbial processes will most likely dominate. 2.4 Ionic mercury [Hg(II)] reduction Reduction of Hg(II) to Hg(0), which occurs in natural waters [83] and soils [84, 85], results in partitioning of Hg into the air due to Hg(0)'s low aqueous solubility (60 μg l−1 water at 25°C) and high volatility (Henry's coefficient of 0.3) [39]. Hg in rain and snow (the major means for the global redistribution of Hg) is largely in the ionic form [86] and is highly available for reduction. Reduction of the deposited Hg(II) to Hg(0) transports Hg back to the atmosphere [87] and therefore prevents its precipitation and settling to bottom sediments where it can be methylated [51, 88]. Indeed, biological reduction contributes significantly to the flux of Hg from natural waters into the atmosphere. Nutrient-rich waters near the equatorial upwelling in the Pacific Ocean are supersaturated with Hg(0) [89] and, in latitudinal transects, Hg(0) concentration correlated positively with biological productivity [90]. Filtration [91] or autoclaving [92] inhibits Hg volatilization from natural waters. Environments highly contaminated with Hg(II) enrich for populations of resistant bacteria [92, 93] and induce mer operon-mediated reduction [94–96]. However, Hg concentrations in most natural environments, including those where MeHg bioaccumulation occurs, are in the fM to pM range. Although mer-lux transcriptional fusions can be induced at 9. Mechanistic studies revealed retention of the skeletal configuration of the substrate consistent with the rare SE2 mechanism rather than a radical-based mechanism. Deuterium isotope effects were consistent with a rate-limiting proton delivery step. Relevant chemical model studies [240] demonstrated a 1000-fold acceleration of aryl-Hg protonolysis by a compound capable of bis-coordination, whereas a monothiol reagent provided only a 50-fold rate acceleration, suggesting that in MerB two protein thiol groups might be involved in stabilizing a reaction intermediate. Recent phylogenetic analysis makes it clear that MerB is a unique enzyme without known homologs which has evolved into several subgroups distinguishable on the basis of their primary and predicted secondary structures [241]. Consistent with its carriage by transferrable plasmids, the MerB phylogeny does not map onto the phylogeny of the bacteria in which those sequenced examples have occurred, i.e. several examples of MerB found in distantly related Gram-positive and Gram-negative bacteria cluster closely and are equidistant from examples found in other Gram-positive and Gram-negative bacteria. Three cysteines are highly conserved at positions 96, 117, and 159 (numbering as for MerB of R831) and consistent with its high requirement for thiols, MerB is a cytosolic enzyme with no disulfide bonds. Cys96 and Cys159 are essential for catalysis and Cys117 appears to have a structural role rather than a catalytic role. MerB works equally well with either of the physiological thiols, glutathione or cysteine. Most described Gram-positive mercury resistance operons are broad-spectrum (i.e. have one or more MerB genes; see Section 4). However, in Gram-negative bacteria the prevalence is closer to 20%[23]. Broad-spectrum-resistance strains of both Gram-positive and Gram-negative bacteria frequently have two mer operons, a broad-spectrum locus and a narrow-spectrum locus, sometimes on the same plasmid as is the case with R831 [242] and pDU1358 [243]. Interestingly, in many occurrences the operon bearing the merB gene in broad-spectrum-resistant Enterobacteriaceae has undergone deletions of varying lengths removing most of the intervening genes between merB and the operonic promoter [23] resulting in the merB gene being much closer to that promoter. In these internally deleted operons, the associated merR gene is retained, possibly because the MerR protein of the co-resident narrow-spectrum resistance operon will not respond to organomercurials [244]. Fig. 7 is an updated reaction model embodying the current state of knowledge concerning MerB. In step 1, a cysteine (probably Cys159 [241]) of the fully reduced enzyme displaces the solvent thiol adduct from the organomercurial and a second protein cysteine (probably Cys96) forms a bis-coordinate structure with the aryl mercurial (step 2). The proton may be donated to this activated bis-coordinated complex by Cys96 or perhaps by some other protonated residue in MerB such as the highly conserved Tyr93. Once the protonated organic moiety leaves, MerB is stuck with Hg(II) (step 3) until two solvent monothiols can remove it (step 4). Interestingly, dithiothreitol (DTT) actually inhibits MerB, possibly by forming a stable three-coordinate complex with the product Hg(II) and one of MerB's cysteines. DTT inhibition can be slowly reversed by either cysteine or glutathione. Figure 7 Open in new tabDownload slide Roles of Thiols in MerB protonolysis of organomercurials. Small red numbers designate positions of cysteine residues in the primary structure of MerB of plasmid R831b [241]. RSH is a low-molecular-mass, cytosolic thiol redox buffer such as glutathione. Ar refers to an aryl (aromatic) moiety in an organomercurial compound such as phenylmercuric acetate. Figure 7 Open in new tabDownload slide Roles of Thiols in MerB protonolysis of organomercurials. Small red numbers designate positions of cysteine residues in the primary structure of MerB of plasmid R831b [241]. RSH is a low-molecular-mass, cytosolic thiol redox buffer such as glutathione. Ar refers to an aryl (aromatic) moiety in an organomercurial compound such as phenylmercuric acetate. While this model accounts for the minimum two-fold excess of thiol required for any catalytic turnover as observed by others, it does not explicitly address the paradoxically high (in vitro) pH optimum of MerB nor the requirement for at least a 15-fold molar excess of thiol for optimum activity [238, 239, 241]. 4 Diversity of mercury resistance loci 4.1 Operon structure The mer locus is widely distributed among eubacterial lineages and mer-like sequences have been identified in several archaean genomes, such as Sulfolobus solfataricus, Thermoplasma volcanicum and Halobacterium sp., though to the best of our knowledge functional mer in Archaea have not been described to date. Several variations on the structure and organization of mer operons are known (Fig. 8) reflecting the mosaic nature of the operon [23, 245]. There seem to be some characteristic differences between the operons of Gram-positive [26, 246–249] and Gram-negative bacteria [161, 243, 250–254]. These differences are: (i) merB is more common in Gram-positive mer operons described to date than in Gram-negative operons; (ii) merR in low-GC Gram-positive operons is transcribed in the same direction as the rest of the operon's genes but in the high-GC Gram-positive Streptomyces operons and all Gram-negative operons merR is transcribed divergently from the structural genes. The Gram-negative marine bacterium Pseudoalteromonas haloplanktis is the exception with merR cotranscribed with merTPCAD[253]. Low homology with other Gram-negative mer genes and additional molecular characteristics further distinguish P. haloplanktis's mer from other Gram-negative operons described to date [253]. The merR-merD-less operon on plasmid pMERPH from Shewanella putrefaciens is another exception to the Gram-negative mer operon pattern though the plasmid does carry an unidentified regulator of merTPCA[254]. Figure 8 Open in new tabDownload slide Diversity of mer operons. Sequenced mer operons from Gram-positive (above line) and Gram-negative (below line) bacteria (see text for citations). Arrows indicate the direction of translation of each gene product. Colorless arrows indicate ORFs with unknown functions. Figure 8 Open in new tabDownload slide Diversity of mer operons. Sequenced mer operons from Gram-positive (above line) and Gram-negative (below line) bacteria (see text for citations). Arrows indicate the direction of translation of each gene product. Colorless arrows indicate ORFs with unknown functions. 4.2 mer as a part of mobile genetic elements The mercury resistance (mer) operon is the metal resistance locus whose dissemination by lateral gene transfer is best established [23, 25, 245, 255]. One archetypal mer operon is carried by Tn21, a composite transposon originally isolated on plasmid NR1 from Shigella flexneri in Japan [256]. Tn21 is a Tn3-like class II replicative transposon that carries genes for a transposase (tnpA), a resolvase (tnpR), a res site for cointegrate resolution, and a class 1 integron, and is flanked by 38-bp inverted repeats [161, 257]. Mercury resistance operons are often part of group II transposons in both Gram-negative [258, 259] and Gram-positive [164, 260] bacteria and these transposons often carry integrons with one or more antibiotic resistance genes [259, 261]. A recent survey of the literature revealed 98 independent examples of mer transposons (J. Coombs, personal communication). Two groups have examined the distribution and diversity of Tn21-type transposons and their associated mer operons in soil bacteria. Strike et al. used probes, polymerase chain reaction primers, restriction fragment length polymorphism (RFLP) and sequence data to show linkage of mer with tnpR, tnpA, and integrons in bacterial isolates from mercury-contaminated and pristine soils [262, 263] and in DNA extracts from the same environments [264]. While most of the isolates carried both group II-type tnpA and tnpR, RFLP patterns indicated a high diversity of these genes [262]. Nikiforov et al., who have characterized mer transposons from soil bacteria by employing a transposon capture approach [265], showed a high occurrence of recombination resulting in chimeric structures of many transposons [258]. For example, recent, exciting work on mer loci carried by the widely distributed Tn5041 transposon in pseudomonads from soil, freshwater, and mercury mine tailings [245] has demonstrated the high level of recombinational promiscuity in this locus and in such elements in general. The basic Tn5041 mer operon has merRTPCAD in an arrangement like that of mer in Tn21 (Figs. 1 and 8). Variants of Tn5041 mer have suffered transposition of the entire Tn21 mer locus (lacking In2 integron) into the 5′ region adjacent to Tn5041 merA. Other Tn5041 mer variants have experienced insertion of an extremely variegated locus called mer2, which encodes merRTPAGBD. This event appears to have arisen by Chi-mediated homologous recombination of a mer operon on a covalently closed circle that was generated by homologous recombination at its flanking IS1015 sequences. Such mosaic structures as recognized by Kholodii et al. [245] also exist in previously sequenced mer loci recovered from feces of monkeys exposed to very high levels of Hg(II) released from amalgam dental restorations [23]. Thus, mer loci reveal variation in structure of the operon and its individual genes and evolution by frequent horizontal gene transfer and recombination events [157]. It is likely that the use of antibiotics in the clinical setting and environmental Hg contamination exert selective pressure that fosters the persistence of mer and its associated mobile elements. Consistent with this idea is the recent finding that mer transposons in clinical isolates from the preantibiotic era had >99.9% DNA sequence identity with Tn1696, Tn5036, Tn5053, and Tn21[266]. 5 Applied biology of mercury resistance The exploitation of mer-mediated functions for the removal of mercury from wastewater was proposed and demonstrated nearly 20 years ago [267]. By now several approaches, employing engineered and naturally occurring mer operons, have been constructed and their application demonstrated in a variety of contaminated industrial streams and environments. The most advanced of these, and the only one in industrial use, is a packed bed bioreactor inoculated with a mixture of several natural isolates of mercury-resistant bacteria (mostly pseudomonads) for the treatment of chlor-alkali process wastewater [268]. The inoculated strains grow during the operation of the bioreactor to form a thick biofilm of cells embedded in exopolysaccharides (EPS) that fill the cavities of the pumice granules that constitute the packed bed [269]. Because the feed wastewater is not sterile invading indigenous mercury-resistant strains can also become established within the bioreactor [269–271]. The bioreactor is fed with a neutralized wastewater supplemented with low concentrations of sucrose and yeast extract, at a flow rate of up to 2000 l h−1 of wastewater containing up to 10 mg l−1 mercury. Stable operation of up to 240 days with 99% efficiency in mercury removal has been reported [269]. Residual mercury in bioreactor effluents is removed in an activated charcoal filter where both adsorption and further microbial reduction contribute to an additional six-fold decrease in effluent mercury concentrations to below 50 μg l−1, the regulated level of Hg in wastewater streams [268, 269]. EPS may limit diffusion in the reactor, resulting in accumulation of reduced Hg as droplets of elemental Hg(0) within the bed matrix [270]. This strategy can afford the safe disposal of the reduced mercury or, if commercially feasible, subsequent recycling. The long-term operation of the industrial-scale bioreactor is challenged by fluctuations in influent mercury concentrations and variations in other wastewater parameters. Von Canstein et al. [269, 272] recently showed that stable bioreactor operation under varying conditions is facilitated by the diversity of the mercury-resistant community. Thus, at high mercury loads the zone of active reduction migrated up the bottom-to-top bioreactor where only a few and most resistant members of the community were present while at low mercury loads the active zone migrated down the bioreactor. Strains with lower levels of mercury resistance were abundant downstream of the reduction front and in the activated charcoal filter [269]. Experimental bioreactors inoculated with a single resistant strain and fed with sterile chlor-alkali wastewater collapsed when influent mercury concentrations increased to 10 mg l−1, whereas a stable performance was achieved by a mixture of the strains [272]. The in situ bioremediation of mercury is more challenging than end-of-pipe treatment, because mercury persisting in the environment is associated with complex matrices and is usually in sulfidic forms [273]. Note that while HgS is highly insoluble and is considered an inert form of mercury, the formation of Hg(0) from HgS was reported for a Thiobacillus ferrooxidans strain that was isolated from a pyrite mine [274]. Organisms that carry complete and partial mer operons have been prepared and tested with the goal of remediating mercury-contaminated soil, sediment, and subsurface environments. An increased rate of Hg(0) evolution from microcosms simulating a contaminated pond by the application of active mercury-reducing strains (bioaugmentation) was reported [275]. The highly radiation-resistant strain Deinococcus radiodurans, to which Tn21 was transferred, reduced Hg(II) in the presence of 6000 rad h−1[276]. Such strains were constructed with the prospect of remediating US Department of Energy subsurface sites containing radionuclides, metals, and organic contaminants [277]. Constitutive mer operons were prepared [278] with the prospect of improving wastewater treatment operations [279], but when tested they proved inferior to naturally isolated mercury-reducing strains [272]. The construction of recombinant E. coli with enhanced mercury accumulation due to combined merTP and cytoplasmic eukaryotic metallothioneins was reported [280]. Induced cells in hollow-fiber bioreactor retained 99% of the mercury from wastewater containing >2 mg l−1 mercury, various other ions, and at an alkaline pH. Affinity for Hg(II) exceeded that of chelating agents suggesting the feasibility of a combined treatment whereby metals released by chelators are subsequently sequestered by the recombinant strain [281, 282]. Finally, the mer system was used to engineer several species of plants for phytoremediation of inorganic [283, 284] and organomercury [285, 286]. In a recent development, the efficiency of Hg(0) evolution from hydrophobic organomercury was improved 10–70-fold by targeting recombinant MerB to the endoplasmic reticulum or the cell wall in Arabidopsis compared to activities of plants carrying cytoplasmic MerB [287]. However, the engineered plants emit Hg(0) into the air, a matter of concern since the atmospheric deposition of Hg(II), following the oxidation of Hg(0) in the atmosphere, remains a major source of Hg in the environment [40]. Although calculations suggest that plant-enhanced emissions would have little effect on the global atmospheric Hg(0) pool [283], public perceptions have slowed the implementation of this promising mer-based phytoremediation strategy. Lastly, there is considerable interest in using MerR as a biosensor in vivo with various transcriptional fusions [149, 174, 288, 289] and in vitro to optical detection devices [145, 146]. The value of such sensors is in distinguishing bioavailable from total Hg(II). Such tools are needed to improve our understanding of the mercury geochemical cycle and to facilitate regulatory action aimed at controlling human and wildlife exposures in contaminated environments [290]. Such constructs have been valuable for research on extra- and intracellular interactions of Hg(II) with bacterial cells [97, 291], and on the environmental behavior of mercury [145, 146, 174, 289] but, to date, have made only limited contributions to the analysis of bioavailable mercury in environmental [110, 290] and industrial applications. 6 Future directions Despite being the longest studied of the bacterial toxic metal resistance loci, mer continues to bring new insights to gene regulation and enzymology. There are also some large mysteries yet to be solved in the subject of Hg(II) transport as well as more discrete puzzles about how the various mer proteins tune their thiol groups to hold mercurial compounds in just the right way and just long enough to convert them into a less toxic and readily dissipated form. The population biology of ubiquitous mer loci is proving a rich resource for dissecting lateral gene transfer. Study of this bacterial operon also provides insights into a much larger subject: ‘the biology of mercury’. This apparent oxymoron is apt because mercury's protean chemical properties have made it a valuable element, exploited by humans for millennia. Although its toxicity has long been known, mercury's chemical utility allowed humans to overlook its dangers. In the later half of the last century, discovery that organisms as tough as bacteria had evolved a specialized system to deal with Hg toxicity contributed, along with several high profile anthropogenic human exposure disasters, to increased concern for the dangers of Hg to more delicate creatures such as ourselves [292]. Since Hg(II) mimics reactive oxygen species (ROS) in its interactions with cellular constituents, the chronic nature of Hg intoxication in higher animals owes a lot to the fact that eukaryotes have evolved a wealth of strategies for dealing with ROS over the last 2.5 billion years. Recent revelations of the subtlety of the redox-sensing regulators OxyR and SoxR (a MerR family member) [293–295] suggest that examination of Hg(II) interactions with ROS-sensing and redox homeostasis systems in both prokaryotic and eukaryotic cells will be a timely and fruitful pursuit. Acknowledgements We are grateful to Jarek Letowski and Jeffra Schaefer for Figures 1 and 8, respectively. We also very much appreciate the comments of Lena Sahlman and two anonymous reviewers on the manuscript and the support of work in our labs by NIH, NSF, EPA, and DOE. References [1] Moore B. ( 1960 ) A new screen test and selective medium for the rapid detection of epidemic strains of Staphylococcus aureus . Lancet ii , 453 – 458 . Google Scholar Crossref Search ADS WorldCat [2] Richmond M.H. John M. ( 1964 ) Co-transduction by a staphylococcal phage of the genes responsible for penicillinase synthesis and resistance to mercury salts . Nature 202 , 1360 – 1361 . Google Scholar Crossref Search ADS PubMed WorldCat [3] Novick R.P. Roth C. ( 1968 ) Plasmid resistance to inorganic salts in Staphylococcus aureus . J. Bacteriol . 95 , 1335 – 1342 . Google Scholar PubMed OpenURL Placeholder Text WorldCat [4] Smith D.H. ( 1967 ) R factors mediate resistances to mercury, nickel, and cobalt . Science 156 , 1114 – 1116 . Google Scholar Crossref Search ADS PubMed WorldCat [5] Fagerstrom T. Jernelov A. ( 1971 ) Formation of methyl mercury from pure mercuric sulphide in aerobic organic sediment . Water Res . 5 , 121 – 122 . Google Scholar Crossref Search ADS WorldCat [6] Fagerstrom T. Jernelov A. ( 1972 ) Some aspects of the quantitative ecology of mercury . Water Res . 6 , 1139 – 1202 . Google Scholar Crossref Search ADS WorldCat [7] Tonomura K. Kanzaki F. ( 1969 ) The reductive decomposition of organic mercurials by cell-free extracts of a mercury resistant pseudomonad . Biochim. Biophys. Acta 184 , 227 – 229 . Google Scholar Crossref Search ADS PubMed WorldCat [8] Furukawa K. Suzuki T. Tonomura K. ( 1969 ) Decomposition of organic mercurial compounds by mercury-resistant bacteria . Agric. Biol. Chem . 33 , 128 – 130 . Google Scholar Crossref Search ADS WorldCat [9] Tonomura K. Maeda K. Futai F. Nakagami T. Yamada M. ( 1968 ) Stimulative vaporization of phenylmercuric acetate by mercury-resistant bacteria . Nature 217 , 644 – 646 . Google Scholar Crossref Search ADS PubMed WorldCat [10] Schottel J. Mandal A. Clark D. Silver S. Hedges R.W. ( 1974 ) Volatilisation of mercury and organomercurials determined by inducible R-factor systems in enteric bacteria . Nature 251 , 335 – 337 . Google Scholar Crossref Search ADS PubMed WorldCat [11] Clark D.L. Weiss A.A. Silver S. ( 1977 ) Mercury and organomercurial resistances determined by plasmids in Pseudomonas . J. Bacteriol . 132 , 186 – 196 . Google Scholar PubMed OpenURL Placeholder Text WorldCat [12] Summers A.O. Sugarman L.I. ( 1974 ) Cell-free mercury(II)-reducing activity in a plasmid-bearing strain of Escherichia coli . J. Bacteriol . 119 , 242 – 249 . Google Scholar PubMed OpenURL Placeholder Text WorldCat [13] Schottel J.L. ( 1978 ) The mercuric and organomercurial detoxifying enzymes from a plasmid-bearing strain of Escherichia coli . J. Biol. Chem . 253 , 4341 – 4349 . Google Scholar PubMed OpenURL Placeholder Text WorldCat [14] Foster T.J. Nakahara H. Weiss A.A. Silver S. ( 1979 ) Transposon A-generated mutations in the mercuric resistance genes of plasmid R100-1 . J. Bacteriol . 140 , 167 – 181 . Google Scholar PubMed OpenURL Placeholder Text WorldCat [15] Nakahara H. Silver S. Miki T. Rownd R.H. ( 1979 ) Hypersensitivity to Hg2+ and hyperbinding activity associated with cloned fragments of the mercurial resistance operon of plasmid NR1 . J. Bacteriol . 140 , 161 – 166 . Google Scholar PubMed OpenURL Placeholder Text WorldCat [16] Jackson W.J. Summers A.O. ( 1982 ) Polypeptides encoded by the mer operon . J. Bacteriol . 149 , 479 – 487 . Google Scholar PubMed OpenURL Placeholder Text WorldCat [17] Jackson W.J. Summers A.O. ( 1982 ) Biochemical characterization of the HgCl2-inducible polypeptides encoded by the mer operon of plasmid R100 . J. Bacteriol . 151 , 962 – 970 . Google Scholar PubMed OpenURL Placeholder Text WorldCat [18] Laddaga R.A. Chu L. Misra T. Silver S. ( 1987 ) Nucleotide sequence and expression of the mercurial-resistance operon from Staphylococcus aureus plasmid pI258 . Proc. Natl. Acad. Sci. USA 84 , 5106 – 5110 . Google Scholar Crossref Search ADS WorldCat [19] Brown N.L. Misra T.K. Winnie J.N. Schmidt A. Seiff M. Silver S. ( 1986 ) The nucleotide sequence of the mercuric resistance operons of plasmid R100 and transposon Tn501: further evidence for mer genes which enhance the activity of the mercuric ion detoxification system . Mol. Gen. Genet . 202 , 143 – 151 . Google Scholar Crossref Search ADS PubMed WorldCat [20] Barrineau P. Gilbert M.P. Jackson W.J. Jones C.S. Summers A.O. Wisdom S. ( 1984 ) The DNA sequence of the mercury resistance operon of the IncFII plasmid NR1 . J. Mol. Appl. Genet . 2 , 601 – 619 . Google Scholar PubMed OpenURL Placeholder Text WorldCat [21] Wang Y. Moore M.J. Levinson H.S. Silver S. Walsh C.T. ( 1989 ) Nucleotide sequence of a chromosomal mercury resistance determinant from a Bacillus sp. with broad spectrum mercury-resistance . J. Bacteriol . 171 , 83 – 92 . Google Scholar Crossref Search ADS PubMed WorldCat [22] Bogdanova E. Mindlin S. Pakrava E. Kocur M. Rouch D. ( 1992 ) Mercuric reductase in enviromental Gram-positive bacteria sensitive to mercury . FEMS Microbiol. Lett . 97 , 95 – 100 . Google Scholar Crossref Search ADS WorldCat [23] Liebert C.A. Wireman J. Smith T. Summers A.O. ( 1997 ) Phylogeny of mercury resistance (mer) operons of Gram-negative bacteria isolated from the fecal flora of primates . Appl. Environ. Microbiol . 63 , 1066 – 1076 . Google Scholar PubMed OpenURL Placeholder Text WorldCat [24] Steele R.A. Opella S.J. ( 1997 ) Structures of the reduced and mercury-bound forms of MerP, the periplasmic protein from the bacterial mercury detoxification system . Biochemistry 36 , 6885 – 6895 . Google Scholar Crossref Search ADS PubMed WorldCat [25] Osborn A.M. Bruce K.D. Strike P. Ritchie D.A. ( 1997 ) Distribution, diversity, and evolution of the bacterial mercury resistance (mer) operon . FEMS Microbiol. Rev . 19 , 239 – 262 . Google Scholar Crossref Search ADS PubMed WorldCat [26] Huang C.-C. Narita M. Yamagata T. Endo G. ( 1999 ) Identification of three merB genes and characterization of a broad-spectrum mercury resistance module encoded by a class II transposon of Bacillus megaterium MB1 . Gene 239 , 361 – 366 . Google Scholar Crossref Search ADS PubMed WorldCat [27] Shiratori T. Inoue C. Sugawara K. Kusano T. Kitagawa Y. ( 1989 ) Cloning and expression of Thiobacillus ferroxidans mercury ion resistance genes in Escherichia coli . J. Bacteriol . 171 , 3458 – 3464 . Google Scholar Crossref Search ADS PubMed WorldCat [28] Goldstein C. Lee M.D. Sanchez S. Hudson C. Phillips B. Register B. Grady M. Liebert C. Summers A.O. White D.G. Maurer J.J. ( 2001 ) Incidence of class 1 and 2 integrases in clinical and commensal bacteria from livestock, companion animals, and exotics . Antimicrob. Agents Chemother . 45 , 723 – 726 . Google Scholar Crossref Search ADS PubMed WorldCat [29] Bruce K.D. Hiorns W.D. Hobman J.L. Osborn A.M. Strike P. Ritchie D.A. ( 1992 ) Amplification of DNA from native populations of soil bacteria by using polymerase chain reaction . Appl. Environ. Microbiol . 58 , 3413 – 3416 . Google Scholar PubMed OpenURL Placeholder Text WorldCat [30] Osborn A. Bruce K. Strike P. Ritchie D. ( 1993 ) Polymerase chain reaction-restriction fragment length polymorphism analysis shows divergence among mer determinants from Gram-negative soil bacteria indistinguishable by DNA-DNA hybridization . Appl. Environ. Microbiol . 59 , 4024 – 4030 . Google Scholar PubMed OpenURL Placeholder Text WorldCat [31] Hobman J. Kholodii G. Nikiforov V. Ritchie D.A. Strike P. Yurieva O. ( 1994 ) The sequence of the mer operon of pMER327/419 and transposon ends of pMER327/419, 330 and 05 . Gene 277 , 73 – 78 . Google Scholar Crossref Search ADS WorldCat [32] Kholodi G.Y. Gorlenko Z.M. Lomoskaya O.L. Mindlin S.Z. Yurieva O.V. Nikiforov V.G. ( 1993 ) Molecular characterization of an aberrant mercury resistance transposable element from an environmental Acinetobacter strain . Plasmid 30 , 303 – 308 . Google Scholar Crossref Search ADS PubMed WorldCat [33] Kholodii G.Y. Mindlin S.Z. Bass I.A. Yurieva O.V. Minakhina S.V. Nikiforov V.G. ( 1995 ) Four genes, two ends, and a res region are involved in transposition of Tn5053: a paradigm for a novel family of transposons carrying either a mer operon or an integron . Mol. Microbiol . 17 , 1189 – 1200 . Google Scholar Crossref Search ADS PubMed WorldCat [34] Yurieva O. Kholodii G. Minakhin L. Gorlenko J. Kalyaeva E. Mindlin S. Nikiforov V. ( 1997 ) Intercontinental spread of promiscuous mercury-resistance transposons in environmental bacteria . Mol. Microbiol . 24 , 321 – 329 . Google Scholar Crossref Search ADS PubMed WorldCat [35] Kholodii G.Y. Yurieva O.V. Gorlenko Z.M. Mindlin S.Z. Bass I.A. Lomovskya O.L. Kopteva A.V. Nikiforov V.G. ( 1997 ) Tn5041: a chimeric mercury resistance transposon closely related to a toluene degradative transposon Tn4651 . Microbiology 143 , 2549 – 2556 . Google Scholar Crossref Search ADS PubMed WorldCat [36] Bogdanova E. Minakhin L. Bass I.A. Volodin A. Hobman J.L. Nikiforov V.G. ( 2001 ) Class II broad-spectrum mercury resistance transposons in Gram-positive bacteria from natural environments . Res. Microbiol . 152 , 503 – 514 . Google Scholar Crossref Search ADS PubMed WorldCat [37] Wedepohl K.K. ( 1995 ) The composition of the continental crust . Geochim. Cosmochim. Acta 59 , 1217 – 1232 . Google Scholar Crossref Search ADS WorldCat [38] Barnes, H.L. and Seward, T.M. (1997) Geothermal systems and mercury deposits. In: Geochemistry of Hydrothermal Ore Deposits, 3rd edn. (Barnes, H.L., Ed.), pp. 699–736. John Wiley and Sons, New York. [39] Weast, R.C. (1973) Handbook of Chemistry and Physics. CRC Press, Cleveland, OH. [40] National Research Council (2000) Toxicological Effects of Methylmercury. National Academy Press, Washington, DC. [41] Mason, R. and Fitzgerald, W. (1996) The global mercury cycle: Oceanic and anthropogenic aspects. In: Global and Regional Mercury Cycles: Sources, Fluxes and Mass Balances (Baeyens, W., Ed.), pp. 85–108. Kluwer Academic, Dordrecht. [42] Slemer F. Langer E. ( 1992 ) Increase in global atmospheric concentrtions of mercury inferred from measurements over the Atlantic Ocean . Nature 355 , 434 – 437 . Google Scholar Crossref Search ADS WorldCat [43] Fitzgerald W.D. Engstrom R. Mason R. E N. ( 1997 ) The case for atmospheric mercury contamination in remote areas . Environ. Sci. Technol . 32 , 1 – 7 . Google Scholar Crossref Search ADS WorldCat [44] Lorey P. Driscoll C. ( 1999 ) Historical trends of mercury deposition in Adirondack Lakes . Environ. Sci. Technol . 33 , 718 – 722 . Google Scholar Crossref Search ADS WorldCat [45] Swain E.B. Engstrom D. Brigham M. Henning T. Brezonik P. ( 1992 ) Increasing rates of atmospheric mercury deposition in mid-continental North America . Science 257 , 784 – 787 . Google Scholar Crossref Search ADS PubMed WorldCat [46] Schuster P.F. Krabbenhoft D.P. Naftz D.L. Cecil L.D. Olson M.L. Dewild J.F. Susong D.D. Green J.R. Abbott M.L. ( 2002 ) Atmospherc mercury deposition during the last 270 years: a glacial ice core record of natural and anthropogenic sources . Environ. Sci. Technol . 36 , 2303 – 2310 . Google Scholar Crossref Search ADS PubMed WorldCat [47] Morel F.M.M. Kraepiel A.M.L. Amyot M. ( 1998 ) The chemical cycle and bioaccumulation of mercury . Annu. Rev. Ecol. Syst . 29 , 543 – 566 . Google Scholar Crossref Search ADS WorldCat [48] Clarkson T.W. ( 1997 ) The toxicology of mercury . Crit. Rev. Clin. Lab. Sci . 34 , 369 – 403 . Google Scholar Crossref Search ADS PubMed WorldCat [49] Clarkson T.W. ( 2002 ) The three modern faces of mercury . Environ. Health Perspect . 110 ( Suppl. 1 ), 11 – 23 . Google Scholar Crossref Search ADS PubMed WorldCat [50] Lawson N.M. Mason R.P. ( 2001 ) Concentration of mercury, methylmercury, cadmium, lead, arsenic, and selenium in the rain and stream water of two contrasting watersheds in western Maryland . Water Res . 35 , 4039 – 4052 . Google Scholar Crossref Search ADS PubMed WorldCat [51] Barkay T. Turner R. Saouter E. Horn J. ( 1992 ) Mercury biotransformations and their potential for remediation of mercury contamination . Biodegradation 3 , 147 – 159 . Google Scholar Crossref Search ADS WorldCat [52] Jensen S. Jernelöv A. ( 1969 ) Biological methylation of mercury in aquatic organisms . Nature 223 , 753 – 754 . Google Scholar Crossref Search ADS PubMed WorldCat [53] Wood J.M. ( 1974 ) Biological cycles for toxic elements in the environment . Science 183 , 1049 – 1052 . Google Scholar Crossref Search ADS PubMed WorldCat [54] Wood J.M. Scott Kennedy F. Rosen C.G. ( 1968 ) Synthesis of methyl-mercury compounds by extracts of a methanogenic bacterium . Nature 220 , 173 – 174 . Google Scholar Crossref Search ADS PubMed WorldCat [55] Vonk J.W. Sijpesteijn A.K. ( 1973 ) Studies on the methylation of mercuric chlorid by pure cultures of bacteria and fungi . Antonie van Leeuwenhoek 39 , 505 – 513 . Google Scholar Crossref Search ADS PubMed WorldCat [56] Compeau G.C. Bartha R. ( 1985 ) Sulfate-reducing bacteria: principle methylators of mercury in anoxic estuarine sediment . Appl. Environ. Microbiol . 50 , 498 – 502 . Google Scholar PubMed OpenURL Placeholder Text WorldCat [57] Gilmour C.C. Henry E.A. Mitchell R. ( 1992 ) Sulfate stimulation of mercury methylation in freshwater sediments . Environ. Sci. Technol . 26 , 2281 – 2287 . Google Scholar Crossref Search ADS WorldCat [58] King J.K. Kostka J.E. Frischer M.E. Saunders F.M. ( 2000 ) Sulfate-reducing bacteria methylate mercury at variable rates in pure culture and in marine sediments . Appl. Environ. Microbiol . 66 , 2430 – 2437 . Google Scholar Crossref Search ADS PubMed WorldCat [59] Benoit J.M. Gilmour C.C. Mason R.P. ( 2001 ) The influence of sulfide on solid-phase mercury bioavailability for methylation by pure cultures of Desulfobulbus propionicus (1pr3) . Environ. Sci. Technol . 35 , 127 – 132 . Google Scholar Crossref Search ADS PubMed WorldCat [60] Benoit J.M. Gilmour C.C. Mason R.P. ( 2001 ) Aspects of bioavailability of mercury for methylation in pure cultures of Desulfobulbus propionicus (1pr3) . Appl. Environ. Microbiol . 67 , 51 – 58 . Google Scholar Crossref Search ADS PubMed WorldCat [61] Benoit J.M. Gilmour C.C. Mason R. Hayes A. ( 1999 ) Sulfide controls on mercury speciation and bioavailability to methylating bacteria in sediment pore waters . Environ. Sci. Technol . 33 , 951 – 957 . Google Scholar Crossref Search ADS WorldCat [62] Berman M. Chase J.T. Bartha R. ( 1990 ) Carbon flow in mercury biomethylation by Desulfovibrio desulfuricans . Appl. Environ. Microbiol . 56 , 298 – 300 . Google Scholar PubMed OpenURL Placeholder Text WorldCat [63] Choi S.C. Chase J.T. Bartha R. ( 1994 ) Metabolic pathways leading to mercury methylation in Desulfovibrio desulfuricans LS . Appl. Environ. Microbiol . 60 , 4072 – 4077 . Google Scholar PubMed OpenURL Placeholder Text WorldCat [64] Choi S.C. Bartha R. ( 1993 ) Cobalamin-mediated mercury methylation by Desulfovibrio desulfuricans LS . Appl. Environ. Microbiol . 59 , 290 – 295 . Google Scholar PubMed OpenURL Placeholder Text WorldCat [65] Choi S.C. Chase T. Jr. Bartha R. ( 1994 ) Enzymatic catalysis of mercury methylation by Desulfovibrio desulfuricans LS . Appl. Environ. Microbiol . 60 , 1342 – 1346 . Google Scholar PubMed OpenURL Placeholder Text WorldCat [66] Siciliano S.D. Lean D.R. ( 2002 ) Methyltransferase: an enzyme assay for microbial methylmercury formation in acidic soils and sediments . Environ. Toxicol. Chem . 21 , 1184 – 1190 . Google Scholar Crossref Search ADS PubMed WorldCat [67] Rowland I.R. Davies M.J. Grasso P. ( 1977 ) Volatilization of methylmercuric chloride by hydrogen sulphide . Nature 265 , 718 – 719 . Google Scholar Crossref Search ADS PubMed WorldCat [68] Deacon, G.B. Volatilisation of methylmercuric chloride by hydrogen sulphide. Nature. 275, 1978. 344 [69] Baldi F. Pepi M. Filippelli M. ( 1993 ) Methylmercury resistance in Desulovibrio desulfuricans strains in relation to methylmercury degradation . Appl. Environ. Microbiol . 59 , 2479 – 2485 . Google Scholar PubMed OpenURL Placeholder Text WorldCat [70] Mason R.P. Sullivan K.A. ( 1999 ) The distribution and speciation of mercury in the South and Equatorial Atlantic: Biogeochemical cycling on trace substances in the Atlantic Ocean; results from the Intergovernmental Oceanographic Commission baseline surveys in August 1993 and May/June 1996 . Deep-Sea Res. II 46 , 937 – 956 . Google Scholar Crossref Search ADS WorldCat [71] Wallschleager D. Hintelmann H. Evans R.D. Wilken R.D. ( 1995 ) Volatilization of dimethylemrcury and elemental mercury from River Elba floodplain soils . Water Air Soil Pollut . 80 , 1325 – 1329 . Google Scholar Crossref Search ADS WorldCat [72] Weber J.H. ( 1993 ) Review of possible paths for abiotic methylation of mercury(II) in the aquatic environment . Chemosphere 26 , 2063 – 2077 . Google Scholar Crossref Search ADS WorldCat [73] Falter R. ( 1999 ) Experimental study on the unintentional abiotic methylation of inorganic mercury during analysis: part 1: localisation of the compounds effecting the abiotic mercury methylation . Chemosphere 39 , 1051 – 1073 . Google Scholar Crossref Search ADS WorldCat [74] Cerrati G. Bernhard M. Weber J.H. ( 1992 ) Model reactions for abiotic mercury(II) methylation: kinetics of methylation of mercury(II) by mono-, di-, and tri-methylation in seawater . Appl. Organomet. Chem . 6 , 587 – 595 . Google Scholar Crossref Search ADS WorldCat [75] Oremland R.S. Culbertson C.W. Winfrey M.R. ( 1991 ) Methylmercury decomposition in sediments and bacterial cultures: involvement of methanogens and sulfate reducers in oxidative demethylation . Appl. Environ. Microbiol . 57 , 130 – 137 . Google Scholar PubMed OpenURL Placeholder Text WorldCat [76] Marvin-Diapsquale M.C. Oremland R.S. ( 1998 ) Bacterial methylmercury degradation in Florida Everglades peat sediment . Environ. Sci. Technol . 32 , 2556 – 2563 . Google Scholar Crossref Search ADS WorldCat [77] Schaefer J. Latowski J. Barkay T. ( 2002 ) mer-mediated resistance and volatilization of Hg(II) under anaerobic conditions . Geomicrobiol. J . 19 , 87 – 102 . Google Scholar Crossref Search ADS WorldCat [78] Marvin-Diapsquale M.C. Agee J. McGowan C. Oremland R.S. Thomas M. Krabbenhoft D. Gilmour C.C. ( 2000 ) Methyl-mercury degradation pathways: a comparison among three mercury-impacted ecosystems . Environ. Sci. Technol . 34 , 4908 – 4917 . Google Scholar Crossref Search ADS WorldCat [79] Hines M.E. Horvat M. Faganeli J. Bonzongo J.C. Barkay T. Major E.B. Scott K.J. Bailey E.A. Warwick J.J. Lyons W.B. ( 2000 ) Mercury biogeochemistry in the Idrija river, Slovenia, from above the mine into the Gulf of Trieste . Environ. Res . 83 , 129 – 139 . Google Scholar Crossref Search ADS PubMed WorldCat [80] Reniero D. Galli E. Barbieri P. ( 1995 ) Cloning and comparison of mercury- and organomercurial-resistance determinanats from Pseudomonas stutzeri plasmid . Gene 166 , 77 – 82 . Google Scholar Crossref Search ADS PubMed WorldCat [81] Suda I. Suda M. Hirayama K. ( 1993 ) Degradation of methyl and ethyl mercury by singlet oxygen generated from sea water exposed to sunlight or ultraviolet light . Arch. Toxicol . 67 , 365 – 368 . Google Scholar Crossref Search ADS PubMed WorldCat [82] Sellers P. Kelly C.A. Rudd J.W.M. MacHutchon A.R. ( 1996 ) Photodegradation of methylmercury in lakes . Nature 380 , 694 – 697 . Google Scholar Crossref Search ADS WorldCat [83] Nriagu J.O. ( 1994 ) Mechanistic steps in the photoreduction of mercury in natural waters . Sci. Total Environ . 154 , 1 – 8 . Google Scholar Crossref Search ADS WorldCat [84] Schlüter K. ( 2000 ) Review: evaporation of mercury from soils. An integration and synthesis of current knowledge . Environ. Geol . 39 , 249 – 271 . Google Scholar Crossref Search ADS WorldCat [85] Carpi A. Lindberg S.E. ( 1997 ) Sunlight-mediated emission of elemental mercury from soil amended with municipal sewage . Environ. Sci. Technol . 31 , 2085 – 2091 . Google Scholar Crossref Search ADS WorldCat [86] Vandal G.M. Mason R. Fitzgerald W.F. ( 1991 ) Cycling of volatile mercury in temperate lakes . Water Air Soil Pollut . 56 , 791 – 803 . Google Scholar Crossref Search ADS WorldCat [87] Lalonde J.D. Poulain A.J. Amyot M. ( 2002 ) The role of mercury redox reactions in snow on snow-to-air mercury transfer . Environ. Sci. Technol . 36 , 174 – 178 . Google Scholar Crossref Search ADS PubMed WorldCat [88] Fitzgerald W.F. Mason R. Vandal G.M. ( 1991 ) Atmospheric cycling and air-water excahnge of mercury over mid-continental lacusterine regions . Water Air Soil Pollut . 56 , 745 – 767 . Google Scholar Crossref Search ADS WorldCat [89] Fitzgerald W.F. Gill G.A. Kim J.P. ( 1984 ) An equatorial Pacific Ocean source of atmospheric mercury . Science 224 , 597 – 599 . Google Scholar Crossref Search ADS PubMed WorldCat [90] Kim J.P. Fitzgerald W.F. ( 1986 ) Sea-air partitioning of mercury in the equatorial Pacific Ocean . Science 231 , 1131 – 1133 . Google Scholar Crossref Search ADS PubMed WorldCat [91] Sullivan K.A. Mason R.P. ( 1998 ) The concentration and distribution of mercury in Lake Michigan . Sci. Total Environ . 213 , 213 – 228 . Google Scholar Crossref Search ADS WorldCat [92] Barkay T. Turner R. VandenBrook A. Liebert C. ( 1991 ) The relationships of Hg(II) volatilization from a freshwater pond to abundance of mer genes in the gene pool of the indigenous microbial community . Microb. Ecol . 21 , 151 – 161 . Google Scholar Crossref Search ADS PubMed WorldCat [93] Barkay T. ( 1987 ) Adaptation of aquatic microbial communities to Hg2+ stress . Appl. Environ. Microbiol . 53 , 2725 – 2732 . Google Scholar PubMed OpenURL Placeholder Text WorldCat [94] Nazaret S. Jeffrey W.H. Saouter E. von Haven R. Barkay T. ( 1994 ) merA gene expression in aquatic environments measured by mRNA production and Hg(II) volatilization . Appl. Environ. Microbiol . 60 , 4059 – 4065 . Google Scholar PubMed OpenURL Placeholder Text WorldCat [95] Ogunseitan O.A. ( 1998 ) Protein method for investigating mercuric reductase gene expression in aquatic environments . Appl. Environ. Microbiol . 64 , 694 – 702 . OpenURL Placeholder Text WorldCat [96] Siciliano S.D. O'Driscoll N.J. Lean D.R. ( 2002 ) Microbial reduction and oxidation of mercury in freshwater lakes . Environ. Sci. Technol . 36 , 3064 – 3068 . Google Scholar Crossref Search ADS PubMed WorldCat [97] Golding G.R. Kelly C.A. Sparling R. Loewen P.C. Rudd J.W.M. Barkay T. ( 2002 ) Evidence for facilitated uptake of Hg(II) by Vibrio anguillarum and Escherichia coli under anaerobic and aerobic conditions . Limnol. Oceanogr . 47 , 967 – 975 . Google Scholar Crossref Search ADS WorldCat [98] Yu H. Chu L. Misra T.K. ( 1996 ) Intracellular inducer Hg2+ concentration is rate determining for the expression of the mercury-resistance operon in cells . J. Bacteriol . 178 , 2712 – 2714 . Google Scholar Crossref Search ADS PubMed WorldCat [99] Ben-Bassat D. Mayer A.M. ( 1978 ) Light-induced Hg volatilization and O2 evolution in Chlorella and the effct of DCMU and methylamine . Physiol. Plant . 42 , 33 – 38 . Google Scholar Crossref Search ADS WorldCat [100] Devars S. Aviles C. Cervantes C. Moreno-Sanchez R. ( 2000 ) Mercury uptake and removal by Euglena gracilis . Arch. Microbiol . 174 , 175 – 180 . Google Scholar Crossref Search ADS PubMed WorldCat [101] Zhang H. Lindberg S.E. ( 2001 ) Sunlight and iron(III)-induced photochemical production of dissolved gaseous mercury in freshwater . Environ. Sci. Technol . 35 , 928 – 935 . Google Scholar Crossref Search ADS PubMed WorldCat [102] Skogerboe R.K. Wilson S.A. ( 1981 ) Reduction of ionic species by fulvic acid . Anal. Chem . 53 , 228 – 232 . Google Scholar Crossref Search ADS WorldCat [103] Allard B. Arsenie I. ( 1991 ) Abiotic reduction of mercury by humic substances in aquatic system – an important process for the mercury cycle . Water Air Soil Pollut . 56 , 457 – 464 . Google Scholar Crossref Search ADS WorldCat [104] Baltisberger R.J. A H.D. Grieble D. Ballintine T.A. ( 1979 ) A study of the disproportionation of mercury(I) induced by gas sparging in acidic aqueous solutions for cold-vapor atomic absorption spectrometry . Anal. Chim. Acta 111 , 111 – 122 . Google Scholar Crossref Search ADS WorldCat [105] Smith T. Pitts K. McGarvey J.A. Summers A.O. ( 1998 ) Bacterial oxidation of mercury metal vapor, Hg(0) . Appl. Environ. Microbiol . 64 , 1328 – 1332 . Google Scholar PubMed OpenURL Placeholder Text WorldCat [106] Summers A.O. Silver S. ( 1972 ) Mercury resistance in a plasmid-bearing strain of Escherichia coli . J. Bacteriol . 112 , 1228 – 1236 . Google Scholar PubMed OpenURL Placeholder Text WorldCat [107] Hamlett N.V. Landale E.C. Davis B.H. Summers A.O. ( 1992 ) Roles of the Tn21 merT, merP, and merC gene products in mercury resistance and mercury binding . J. Bacteriol . 174 , 6377 – 6385 . Google Scholar Crossref Search ADS PubMed WorldCat [108] Kim, K.-H., Hanson, P.J., Barnett, M.O. and Lindberg, S.E. (1997) Biogeochemistry of mercury in the air-soil-plant system. In: Mercury and its Effects on Environment and Biology, 1st edn., Vol. 34 (Sigel, A. and Sigel, H., Eds), pp. 185–212. Marcel Dekker, New York. [109] Magos L. Clarkson T.W. ( 1978 ) Role of catalase in the oxidation of mercury vapor . Biochem. Pharmacol . 27 , 1373 – 1377 . Google Scholar Crossref Search ADS PubMed WorldCat [110] Lindberg S.E. Brooks S. Lin C.J. Scott K.J. Landis M.S. Stevens R.K. Goodsite M. Richter A. ( 2002 ) Dynamic oxidation of gaseous mercury in the Arctic troposphere at polar sunrise . Environ. Sci. Technol . 36 , 1245 – 1256 . Google Scholar Crossref Search ADS PubMed WorldCat [111] Thoming J. Kliem B.K. Ottosen L.M. ( 2000 ) Electrochemically enhanced oxidation reactions in sandy soil polluted with mercury . Sci. Total Environ . 261 , 137 – 147 . Google Scholar Crossref Search ADS PubMed WorldCat [112] Lalonde J.D. Amyot M. Kraepiel A.M. Morel F.M. ( 2001 ) Photooxidation of Hg(0) in artificial and natural waters . Environ. Sci. Technol . 35 , 1367 – 1372 . Google Scholar Crossref Search ADS PubMed WorldCat [113] Amyot M. Gill G.A. Morel F.M. ( 1997 ) Production and loss of dissolved gaseous mecruy in coastal seawater . Environ. Sci. Technol . 31 , 3606 – 3611 . Google Scholar Crossref Search ADS WorldCat [114] De Magalhaes M.E. Tubino M. ( 1995 ) A possible path for mercury in biological systems: the oxidation of metallic mercury by molecular oxygen in aqueous solutions . Sci. Total Environ . 170 , 229 – 239 . Google Scholar Crossref Search ADS PubMed WorldCat [115] Munthe J. ( 1992 ) The aqueous oxidation of elemental mercury by ozone . Atmos. Environ . 26A , 1461 – 1468 . Google Scholar Crossref Search ADS WorldCat [116] Seigneur C. Wrobel J. Constantinou E. ( 1994 ) A chemical kinetic mechanism for atmospheric inorganic mercury . Environ. Sci. Technol . 28 , 1589 – 1597 . Google Scholar Crossref Search ADS PubMed WorldCat [117] Yamamoto M. ( 1995 ) Possible mechanism of elemental mercury oxidation in the presence of SH compounds in aqueous solution . Chemosphere 31 , 2791 – 2798 . Google Scholar Crossref Search ADS WorldCat [118] Ebinghaus R. Kock H.H. Temme C. Einax J.W. Lowe A.G. Richter A. Burrows J.P. Schroeder W.H. ( 2002 ) Antarctic springtime depletion of atmospheric mercury . Environ. Sci. Technol . 36 , 1238 – 1244 . Google Scholar Crossref Search ADS PubMed WorldCat [119] Brown, N.L., Stoyanov, J.V., Kidd, S.P. and Hobman, J.L. (2003) The MerR family of transcriptional regulators. FEMS Microbiol. Rev., 145–163. [120] Caguiat J. Watson A.L. Summers A.O. ( 1999 ) Cd(II)-responsive and constitutive mutants implicate a novel domain in MerR . J. Bacteriol . 181 , 3462 – 3471 . Google Scholar PubMed OpenURL Placeholder Text WorldCat [121] Park S.-J. Wireman J. Summers A.O. ( 1992 ) Genetic analysis of the Tn21 mer operator-promoter . J. Bacteriol . 174 , 2160 – 2171 . Google Scholar Crossref Search ADS PubMed WorldCat [122] Ross W. Park S.-J. Summers A.O. ( 1989 ) Genetic analysis of transcriptional activation and repression in the Tn21 mer operon . J. Bacteriol . 171 , 4009 – 4018 . Google Scholar Crossref Search ADS PubMed WorldCat [123] Heltzel A. Lee I.W. Totis P.A. Summers A.O. ( 1990 ) Activator-dependent preinduction binding of sigma-70 RNA polymerase at the metal-regulated mer promoter . Biochemistry 29 , 9572 – 9584 . Google Scholar Crossref Search ADS PubMed WorldCat [124] Kulkarni R. Summers A.O. ( 1999 ) MerR crosslinks to the α, β, and σ70 subunits of RNA polymerase in the preinitiation complex at the merTPCAD promoter . Biochemistry 38 , 3362 – 3368 . Google Scholar Crossref Search ADS PubMed WorldCat [125] Ansari A.Z. Chael M.L. O'Halloran T.V. ( 1991 ) Allosteric underwinding of DNA is a critical step in positive control of transcription by Hg-MerR . Nature 355 , 87 – 89 . Google Scholar Crossref Search ADS WorldCat [126] Ansari A.Z. Bradner J.E. O'Halloran T.V. ( 1995 ) DNA-bend modulation in a repressor-to-activator switching mechanism . Nature 374 , 371 – 375 . Google Scholar Crossref Search ADS PubMed WorldCat [127] Lee I.W. Livrelli V. Park S.-J. Totis P.A. Summers A.O. ( 1993 ) In vivo DNA-protein interactions at the divergent mercury resistance (mer) promoters II. Repressor/activator (MerR)-RNA polymerase interaction with merOP mutants . J. Biol. Chem . 268 , 2632 – 2639 . Google Scholar PubMed OpenURL Placeholder Text WorldCat [128] Livrelli V. Lee I.W. Summers A.O. ( 1993 ) In vivo DNA-protein interactions at the divergent mercury resistance (mer) promoters. I Metalloregulatory protein MerR mutants . J. Biol. Chem . 268 , 2623 – 2631 . Google Scholar PubMed OpenURL Placeholder Text WorldCat [129] Ptashne, M. and Gann, A. (2002) Genes and Signals. Cold Spring Harbor Laboratory Press, Cold Spring Harbor, NY. [130] Helmann J.D. Ballard B.T. Walsh C.T. ( 1990 ) The MerR metalloregulatory protein binds mercuric ion as a tricoordinate, metal-bridged dimer . Science 247 , 946 – 948 . Google Scholar Crossref Search ADS PubMed WorldCat [131] Shewchuk L.M. Helmann J.D. Ross W. Park S.J. Summers A.O. Walsh C.T. ( 1989 ) Transcriptional switching by the MerR protein: Activation and repression mutants implicate distinct DNA and mercury(II) binding domains . Biochemistry 28 , 2340 – 2344 . Google Scholar Crossref Search ADS PubMed WorldCat [132] Shewchuk L.M. Verdine G.L. Nash H. Walsh C.T. ( 1989 ) Mutagenesis of the cysteines in the metalloregulatory protein MerR indicates that a metal-bridged dimer via Cys126 activates transcription . Biochemistry 28 , 6140 – 6145 . Google Scholar Crossref Search ADS PubMed WorldCat [133] Shewchuk L.M. Verdine G.L. Walsh C.T. ( 1989 ) Transcriptional switching by the metalloregulatory MerR protein: Initial characterization of DNA and mercury(II) binding activities . Biochemistry 28 , 2331 – 2339 . Google Scholar Crossref Search ADS PubMed WorldCat [134] Zeng Q. Stålhandske C. Anderson M.C. Scott R.A. Summers A.O. ( 1998 ) The core metal recognition domain of MerR . Biochemistry 37 , 15885 – 15895 . Google Scholar Crossref Search ADS PubMed WorldCat [135] Godsey M.H. Baranova N.N. Neyfakh A.A. Brennan R.G. ( 2001 ) Crystal structure of MtaN, a global multidrug transporter gene activator . J. Biol. Chem . 276 , 47178 – 47184 . Google Scholar Crossref Search ADS PubMed WorldCat [136] Heldwein E.E.Z. Brennan R.G. ( 2001 ) Crystal structure of the transcription activator BmrR bound to DNA and a drug . Nature 409 , 378 – 382 . Google Scholar Crossref Search ADS PubMed WorldCat [137] O'Halloran T.V. Walsh C. ( 1987 ) Metalloregulatory DNA-binding protein encoded by the merR gene: isolation and characterization . Science 235 , 211 – 214 . Google Scholar Crossref Search ADS PubMed WorldCat [138] Utschig L.M. Bryson J.W. O'Halloran T.V. ( 1995 ) Mercury-199 NMR of the metal receptor site in MerR and its protein-DNA complex . Science 268 , 380 – 385 . Google Scholar Crossref Search ADS PubMed WorldCat [139] Brocklehurst K.R. Hobman J.L. Lawley B. Blank L. Marshall S.J. Brown N.L. Morby A.P. ( 1998 ) ZntR is a Zn(II)-responsive MerR-like transcriptional regulator of zntA in Escherichia coli . Mol. Microbiol . 31 , 893 – 902 . Google Scholar Crossref Search ADS WorldCat [140] Hitomi Y. Outten C.E. O'Halloran T.V. ( 2001 ) Extreme zinc-binding thermodynamics of the metal sensor/regulator protein, ZntR . J. Am. Chem. Soc . 123 , 8614 – 8615 . Google Scholar Crossref Search ADS PubMed WorldCat [141] Outten C.E. Outten F.W. O'Halloran T.V. ( 1999 ) DNA distortion mechanism for transcriptional activation by ZntR, a Zn(II)-responsive MerR homologue in Escherichia coli . J. Biol. Chem . 274 , 37517 – 37524 . Google Scholar Crossref Search ADS PubMed WorldCat [142] Xiong A. Jayaswal R.K. ( 1998 ) Molecular characterization of a chromosomal determinant conferring resistance to zinc and cobalt ions in Staphylococcus aureus . J. Bacteriol . 180 , 4024 – 4029 . Google Scholar PubMed OpenURL Placeholder Text WorldCat [143] Ralston D.M. O'Halloran T.V. ( 1990 ) Ultrasensitivity and heavy-metal selectivity of the allosterically modulated MerR transcription complex . Proc. Natl. Acad. Sci. USA 87 , 3846 – 3850 . Google Scholar Crossref Search ADS WorldCat [144] Wright, J.G. (1991) Spectroscopic and chemical characterization of the mercury receptor site in the transcriptional activator MerR. Ph.D. Dissertation. Northwestern University, Evanston, IL. [145] Bontidean I. Berggren C. Johansson G. Csoregi E. Mattiasson B. Lloyd J.R. Jakeman K.J. Brown N.L. ( 1998 ) Detection of heavy metal ions at femtomolar levels using protein-based biosensors . Anal. Chem . 70 , 4162 – 4169 . Google Scholar Crossref Search ADS PubMed WorldCat [146] Bontidean I. Lloyd J.R. Hobman J.L. Wilson J.R. Csoregi E. Mattiasson B. Brown N.L. ( 2000 ) Bacterial metal-resistance proteins and their use in biosensors for the detection of bioavailable heavy metals . J. Inorg. Biochem . 79 , 225 – 229 . Google Scholar Crossref Search ADS PubMed WorldCat [147] Parkhill J. Brown N.L. ( 1990 ) Site-specific insertion and deletion mutants in the mer promoter-operator region of Tn501; The 19bp spacer is essential for normal induction of the promoter by MerR . Nucleic Acids Res . 18 , 5157 – 5162 . Google Scholar Crossref Search ADS PubMed WorldCat [148] Frantz B. O'Halloran T.V. ( 1990 ) DNA distortion accompanies transcriptional activation by the metal-responsive gene-regulatory protein MerR . Biochemistry 29 , 4747 – 4751 . Google Scholar Crossref Search ADS PubMed WorldCat [149] Condee C.W. Summers A.O. ( 1992 ) A mer-lux transcriptional fusion for real-time examination of in vivo induction kinetics and promoter response to altered superhelicity . J. Bacteriol . 174 , 8094 – 8101 . Google Scholar Crossref Search ADS PubMed WorldCat [150] Caslake L.F. Ashraf S.I. Summers A.O. ( 1997 ) Mutations in the alpha and sigma-70 subunits of RNA polymerase affect expression of the mer operon . J. Bacteriol . 179 , 1787 – 1795 . Google Scholar Crossref Search ADS PubMed WorldCat [151] O'Halloran T.V. Frantz B. Shin M.K. Ralston D.M. Wright J.G. ( 1989 ) The MerR heavy metal receptor mediates positive activation in a topologically novel transcription complex . Cell 56 , 119 – 129 . Google Scholar Crossref Search ADS PubMed WorldCat [152] Lee I.W. Gambill B.D. Summers A.O. ( 1989 ) Translation of merD in Tn21 . J. Bacteriol . 171 , 2222 – 2225 . Google Scholar Crossref Search ADS PubMed WorldCat [153] Mukhopadhyay D. Yu H. Nucifora G. Misra T.K. ( 1991 ) Purification and functional characterization of MerD: A coregulator of the mercury resistance operon in Gram-negative bacteria . J. Biol. Chem . 266 , 18538 – 18542 . Google Scholar PubMed OpenURL Placeholder Text WorldCat [154] Nucifora G. Silver S. Misra T.K. ( 1990 ) Down regulation of the mercury resistance operon by the most promoter-distal gene merD . Mol. Gen. Genet . 220 , 69 – 72 . OpenURL Placeholder Text WorldCat [155] Gambill B.D. Summers A.O. ( 1992 ) Synthesis and degradation of the mRNA of the Tn21 mer operon . J. Mol. Biol . 225 , 251 – 259 . Google Scholar Crossref Search ADS PubMed WorldCat [156] Hawkins H.C. Freedman R.B. ( 1976 ) Thiol-protein disulphide oxidoreductases . Biochem. J . 159 , 385 – 393 . Google Scholar Crossref Search ADS PubMed WorldCat [157] Liebert C.A. Watson A.L. Summers A.O. ( 2000 ) The quality of merC, a module of the mer mosaic . J. Mol. Evol . 51 , 607 – 622 . Google Scholar Crossref Search ADS PubMed WorldCat [158] Foster T.J. ( 1987 ) The genetics and biochemistry of mercury resistance . CRC Crit. Rev. Microbiol . 15 , 117 – 140 . Google Scholar Crossref Search ADS WorldCat [159] Summers A.O. ( 1986 ) Organization, expression and evolution of genes for mercury resistance . Annu. Rev. Microbiol . 40 , 607 – 634 . Google Scholar Crossref Search ADS PubMed WorldCat [160] Bloomfield, V.A., Crothers, D.M. and Tinoco, I. Jr. (1974) Physical Chemistry of Nucleic Acids. Harper and Row, New York. [161] Liebert C.A. Hall R.M. Summers A.O. ( 1999 ) Tn21: Flagship of the floating genome . Microbiol. Mol. Biol. Rev . 63 , 507 – 522 . Google Scholar PubMed OpenURL Placeholder Text WorldCat [162] Brown N.L. Shih Y.-C. Leang C. Glendinning K.J. Hobman J.L. Wilson J.R. ( 2002 ) Mercury transport and resistance . Biochem. Soc. Trans . 30 , 715 – 718 . Google Scholar Crossref Search ADS PubMed WorldCat [163] Wilson J.R. Leang C. Morby A.P. Hobman J.L. Brown N.L. ( 2000 ) MerF is a mercury transport protein: different structures but a common mechanism for mercuric ion transporters . FEBS Lett . 472 , 78 – 82 . Google Scholar Crossref Search ADS PubMed WorldCat [164] Huang C.-C. Narita M. Yamagata T. Itoh Y. Endo G. ( 1999 ) Structure analysis of a class II transposon encoding the mercury resistance of the Gram-positive bacterium Bacillus megaterium MB1, a strain isolated from Minimata Bay, Japan . Gene 234 , 361 – 369 . Google Scholar Crossref Search ADS PubMed WorldCat [165] Bull P.C. Cox D.W. ( 1994 ) Wilson disease and Menkes disease: new handles on heavy-metal transport . Trends Genet . 10 , 246 – 252 . Google Scholar Crossref Search ADS PubMed WorldCat [166] O'Halloran T.V. Culotta V.C. ( 2000 ) Metallochaperones, an intracellular shuttle service for metal ions . J. Biol. Chem . 275 , 25057 – 26060 . Google Scholar Crossref Search ADS PubMed WorldCat [167] Sahlman L. Jonsson B.-H. ( 1992 ) Purification and properties of the mercuric-ion-binding protein MerP . Eur. J. Biochem . 205 , 375 – 381 . Google Scholar Crossref Search ADS PubMed WorldCat [168] Sahlman L. Skarfstad E.G. ( 1993 ) Mercuric ion binding abilities of MerP variants containing only one cysteine . Biochem. Biophys. Res. Commun . 196 , 583 – 588 . Google Scholar Crossref Search ADS PubMed WorldCat [169] Morby A.P. Hobman J.L. Brown N.L. ( 1995 ) The role of cysteine residues in the transport of mercuric ions by the Tn501 MerT and MerP mercury-resistance proteins . Mol. Microbiol . 17 , 25 – 35 . Google Scholar Crossref Search ADS PubMed WorldCat [170] Qian H. Sahlman L. Eriksson P.-O. Hambraeus C. Edlund U. Sethson I. ( 1998 ) NMR solution structure of the oxidized form of MerP, a mercuric ion binding protein involved in bacterial mercuric ion resistance . Biochemistry 37 , 9316 – 9322 . Google Scholar Crossref Search ADS PubMed WorldCat [171] Powlowski J. Sahlman L. ( 1999 ) Reactivity of the two essential cysteine residues of the periplasmic mercuric ion-binding protein, MerP . J. Biol. Chem . 274 , 33320 – 33326 . Google Scholar Crossref Search ADS PubMed WorldCat [172] Takahashi N. Creighton T.E. ( 1996 ) On the reactivity and ionization of the active site cysteine residues of Escherichia coli thioredoxin . Biochemistry 35 , 8342 – 8353 . Google Scholar Crossref Search ADS PubMed WorldCat [173] Grauschopf U. Winther J.R. Korber P. Zander T. Dallinger P. Bardwell J.C.A. ( 1995 ) Why is DsbA such an oxidizing disulfide catalyst . Cell 83 , 947 – 955 . Google Scholar Crossref Search ADS PubMed WorldCat [174] Rasmussen L. Turner R.R. Barkay T. ( 1997 ) Cell-density-dependent sensitivity of a mer-lux bioassay . Appl. Environ. Microbiol . 63 , 3291 – 3293 . Google Scholar PubMed OpenURL Placeholder Text WorldCat [175] Barkay T. Gillman M. Turner R.R. ( 1997 ) Effects of dissolved organic carbon and salinity on bioavailability of mercury . Appl. Environ. Microbiol . 63 , 4267 – 4271 . Google Scholar PubMed OpenURL Placeholder Text WorldCat [176] DeSilva T. Veglia G. Porcelli F. Pranter A.M. Opella S.J. ( 2002 ) Selectivity in heavy metal-binding peptides and proteins . Biopolymers 64 , 189 – 197 . Google Scholar Crossref Search ADS PubMed WorldCat [177] Lund P.A. Brown N.L. ( 1987 ) Role of the merT and merP gene products of transposon Tn501 in the induction and expression of resistance to mercuric ions . Gene 52 , 207 – 214 . Google Scholar Crossref Search ADS PubMed WorldCat [178] Sahlman L. Wong W. Powlowski J. ( 1997 ) A mercuric ion uptake role for the integral inner membrane protein, MerC, involved in bacterial mercuric ion resistance . J. Biol. Chem . 272 , 29518 – 29526 . Google Scholar Crossref Search ADS PubMed WorldCat [179] Wireman J. Liebert C.A. Smith C.T. Summers A.O. ( 1997 ) Association of mercury resistance and antibiotic resistance in the Gram negative fecal bacteria of primates . Appl. Environ. Microbiol . 63 , 4494 – 4503 . Google Scholar PubMed OpenURL Placeholder Text WorldCat [180] Kusano T. Ji G. Inoue C. Silver S. ( 1990 ) Constitutive synthesis of a transport function encoded by the Thiobacillus ferrooxidans merC gene cloned in Escherichia coli . J. Bacteriol . 172 , 2688 – 2692 . Google Scholar Crossref Search ADS PubMed WorldCat [181] Olson B.H. Lester J.N. Cayless S.M. Ford S. ( 1992 ) Distribution of mercury resistance determinants in bacterial communities of river sediments . Water Res . 23 , 1209 – 1217 . Google Scholar Crossref Search ADS WorldCat [182] Kholodii G.Y. Gorlenko Z.M. Mindlin S.Z. Hobman J.L. Nikiforov V.G. ( 2002 ) Tn5041-like transposons: molecular diversity, evolutionary relationships and distribution of distinct variants in environmental bacteria . Microbiology 148 , 3569 – 3582 . Google Scholar Crossref Search ADS PubMed WorldCat [183] Solioz M. Vulpe C. ( 1996 ) CPx-type ATPases: a class of P-type ATPases that pump heavy metals . Trends Biochem. Sci . 21 , 237 – 241 . Google Scholar Crossref Search ADS PubMed WorldCat [184] Uno Y. Kiyono M. Tezuka T. Pan-Hou H. ( 1997 ) Phenylmercury transport mediated by merT-merP genes of Pseudomonas K-62 plasmid pMR26 . Biol. Pharm. Bull . 20 , 107 – 109 . Google Scholar Crossref Search ADS PubMed WorldCat [185] Kiyono M. Omura T. Fujimori H. Pan-Hou H. ( 1995 ) Organomercurial resistance determinants in Pseudomonas K-62 are present on two plasmids . Arch. Microbiol . 163 , 242 – 247 . Google Scholar Crossref Search ADS PubMed WorldCat [186] Kiyono M. Uno Y. Omura T. Pan-Hou H. ( 2000 ) Role of MerT and MerP from Pseudomonas K-62 plasmid pMR26 in the transport of phenylmercury . Biol. Pharm. Bull . 23 , 279 – 282 . Google Scholar Crossref Search ADS PubMed WorldCat [187] Kiyono M. Pan-Hou H. ( 1999 ) The merG gene product is involved in phenylmercury resistance in Pseudomonas strain K-62 . J. Bacteriol . 181 , 726 – 730 . Google Scholar PubMed OpenURL Placeholder Text WorldCat [188] Summers A.O. Kight-Olliff L. Slater C. ( 1982 ) Effect of catabolite repression of the mer operon . J. Bacteriol . 149 , 191 – 197 . Google Scholar PubMed OpenURL Placeholder Text WorldCat [189] Selinofova O.V. Barkay T. ( 1994 ) Role of Na+ in the transport of Hg2+ and induction of the Tn21 mer operon . Appl. Environ. Microbiol . 60 , 3503 – 3507 . Google Scholar PubMed OpenURL Placeholder Text WorldCat [190] Higgins C.F. Dorman C.J. Stirling D.A. Waddell L. Booth I.R. May G. Bremer E. ( 1988 ) A physiological role for DNA supercoiling in the osmotic regulation of gene expression in Salmonella typhimurium and Escherichia coli . Cell 52 , 569 – 584 . Google Scholar Crossref Search ADS PubMed WorldCat [191] Furukawa K. Tonomura K. ( 1972 ) Metallic mercury releasing enzyme in mercury-resistant Pseudomonas . Agric. Biol. Chem . 36 , 217 – 226 . Google Scholar Crossref Search ADS WorldCat [192] Tezuka T. Tonomura K. ( 1976 ) Purification and properties of an enzyme catalyzing the splitting of carbon-mercury linkages from mercury-resistant Pseudomonas K-62 strain. I. Splitting enzyme 1 . J. Biochem. (Tokyo) 80 , 79 – 87 . Google Scholar Crossref Search ADS WorldCat [193] Izaki K. Tashiro Y. Funaba T. ( 1974 ) Mechanism of mercuric chloride resistance in microorganisms. 3. Purification and properties of a mercuric ion reducing enzyme from Escherichia coli bearing R factor . J. Biochem. (Tokyo) 75 , 591 – 599 . Google Scholar Crossref Search ADS WorldCat [194] Schottel J.L. ( 1978 ) The mercuric and organomercurial detoxifying enzymes from a plasmid-bearing strain of Escherichia coli . J. Biol. Chem . 253 , 4341 – 4349 . Google Scholar PubMed OpenURL Placeholder Text WorldCat [195] Fox B. Walsh C.T. ( 1982 ) Mercuric reductase. Purification and characterization of a transposon-encoded flavoprotein containing an oxidation-reduction-active disulfide . J. Biol. Chem . 257 , 2498 – 2503 . Google Scholar PubMed OpenURL Placeholder Text WorldCat [196] Fox B.S. Walsh C.T. ( 1983 ) Mercuric reductase: homology to glutathione reductase and lipoamide dehydrogenase. Iodoacetamide alkylation and sequence of the active site peptide . Biochemistry 22 , 4082 – 4088 . Google Scholar Crossref Search ADS PubMed WorldCat [197] Sahlman L. Lindskog S. ( 1983 ) A stopped-flow study of the reaction between mercuric reductase and NADPH . Biochem. Biophys. Res. Commun . 117 , 231 – 237 . Google Scholar Crossref Search ADS PubMed WorldCat [198] Rinderle S.J. Booth J.E. Williams J.W. ( 1983 ) Mercuric reductase from R-plasmid NR1: characterization and mechanistic study . Biochemistry 22 , 869 – 876 . Google Scholar Crossref Search ADS PubMed WorldCat [199] Williams, C.H. Jr. (1992) In: Chemistry and Biochemistry of Flavoenzymes, Vol. 3 (Müller, F., Ed.), pp. 121–211. CRC Press, Boca Raton, FL. [200] Brown N.L. Ford S.J. Pridmore R.D. Fritzinger D.C. ( 1983 ) Nucleotide sequence of a gene from the Pseudomonas transposon Tn501 encoding mercuric reductase . Biochemistry 22 , 4089 – 4095 . Google Scholar Crossref Search ADS PubMed WorldCat [201] Thieme R. Pai E.F. Schirmer R.H. Schulz G.E. ( 1981 ) Three-dimensional structure of glutathione reductase at 2 Å resolution . J. Mol. Biol . 152 , 763 – 782 . Google Scholar Crossref Search ADS PubMed WorldCat [202] Miller S.M. Ballou D.P. Massey V. Williams C.H. Jr. Walsh C.T. ( 1986 ) Two-electron reduced mercuric reductase binds Hg(II) to the active site dithiol but does not catalyze Hg(II) reduction . J. Biol. Chem . 261 , 8081 – 8084 . Google Scholar PubMed OpenURL Placeholder Text WorldCat [203] Sahlman L. Lambeir A.M. Lindskog S. Dunford H.B. ( 1984 ) The reaction between NADPH and mercuric reductase from Pseudomonas aeruginosa . J. Biol. Chem . 259 , 12403 – 12408 . Google Scholar PubMed OpenURL Placeholder Text WorldCat [204] Sahlman L. Lambeir A.M. Lindskog S. ( 1986 ) Rapid-scan stopped-flow studies of the pH dependence of the reaction between mercuric reductase and NADPH . Eur. J. Biochem . 156 , 479 – 488 . Google Scholar Crossref Search ADS PubMed WorldCat [205] Sandstrom A. Lindskog S. ( 1988 ) Rapid-scan stopped-flow studies of the flavoenzyme mercuric reductase during catalytic turnover . Eur. J. Biochem . 173 , 411 – 415 . Google Scholar Crossref Search ADS PubMed WorldCat [206] Schiering N. Kabsch W. Moore M.J. Distefano M.D. Walsh C.T. Pai E.F. ( 1991 ) Structure of the detoxification catalyst mercuric ion reductase from Bacillus sp. strain RC607 . Nature 352 , 168 – 172 . Google Scholar Crossref Search ADS PubMed WorldCat [207] Schultz P.G. Au K.G. Walsh C.T. ( 1985 ) Directed mutagenesis of the redox-active disulfide in the flavoenzyme mercuric ion reductase . Biochemistry 24 , 6840 – 6848 . Google Scholar Crossref Search ADS PubMed WorldCat [208] Distefano M.D. Au K.G. Walsh C.T. ( 1989 ) Mutagenesis of the redox-active disulfide in mercuric ion reductase: catalysis by mutant enzymes restricted to flavin redox chemistry . Biochemistry 28 , 1168 – 1183 . Google Scholar Crossref Search ADS PubMed WorldCat [209] Miller S.M. Massey V. Ballou D. Williams C.H. Jr. Distefano M.D. Moore M.J. Walsh C.T. ( 1990 ) Use of a site-directed triple mutant to trap intermediates: demonstration that the flavin C(4a)-thiol adduct and reduced flavin are kinetically competent intermediates in mercuric ion reductase . Biochemistry 29 , 2831 – 2841 . Google Scholar Crossref Search ADS PubMed WorldCat [210] Moore M.J. Walsh C.T. ( 1989 ) Mutagenesis of the N- and C-terminal cysteine pairs of Tn501 mercuric ion reductase: consequences for bacterial detoxification of mercurials . Biochemistry 28 , 1183 – 1194 . Google Scholar Crossref Search ADS PubMed WorldCat [211] Moore, M., Distefano, M., Walsh, C., Miller, S., Massey, V., Williams, C.H.J. and Ballou, D. (1987) The penultimate cysteines in mercuric reductase aid in the reduction of mercury. In: Flavins and Flavoproteins 1987 (Edmondson, D.E. and McCormick, D.B., Eds.), pp. 41–44. Walter de Gruyter, Berlin. [212] Miller S.M. Moore M.J. Massey V. Williams C.H. Jr. Distefano M.D. Ballou D.P. Walsh C.T. ( 1989 ) Evidence for the participation of Cys558 and Cys559 at the active site of mercuric reductase . Biochemistry 28 , 1194 – 1205 . Google Scholar Crossref Search ADS PubMed WorldCat [213] Sandstrom A. Lindskog S. ( 1987 ) Activation of mercuric reductase by the substrate NADPH . Eur. J. Biochem . 164 , 243 – 249 . Google Scholar Crossref Search ADS PubMed WorldCat [214] Distefano M.D. Moore M.J. Walsh C.T. ( 1990 ) Active site of mercuric reductase resides at the subunit interface and requires Cys135 and Cys140 from one subunit and Cys558 and Cys559 from the adjacent subunit: evidence from in vivo and in vitro heterodimer formation . Biochemistry 29 , 2703 – 2713 . Google Scholar Crossref Search ADS PubMed WorldCat [215] Moore M.J. Miller S.M. Walsh C.T. ( 1992 ) C-terminal cysteines of Tn501 mercuric ion reductase . Biochemistry 31 , 1677 – 1685 . Google Scholar Crossref Search ADS PubMed WorldCat [216] Cotton, F.A., Wilkinson, G., Murillo, C. and Grimes, R.N. (1999) Advanced Inorganic Chemistry, pp. 598–617. John Wiley and Sons, New York. [217] Engst S. Miller S.M. ( 1998 ) Rapid reduction of Hg(II) by mercuric ion reductase does not require the conserved C-terminal cysteine pair using HgBr2 as the substrate . Biochemistry 37 , 11496 – 11507 . Google Scholar Crossref Search ADS PubMed WorldCat [218] Engst S. Miller S.M. ( 1999 ) Alternative routes for entry of HgX2 into the active site of mercuric ion reductase depend on the nature of the X ligands . Biochemistry 38 , 3519 – 3529 . Google Scholar Crossref Search ADS PubMed WorldCat [219] Engst S. Miller S.M. ( 1999 ) Rapid reduction of Hg(II) by mercuric ion reductase does not require the conserved C-terminal cysteine pair using HgBr2 as the substrate – Correction . Biochemistry 38 , 853 – 854 . Google Scholar Crossref Search ADS PubMed WorldCat [220] Rennex D. Pickett M. Bradley M. ( 1994 ) In vivo and in vitro effects of mutagenesis of active site tyrosine residues of mercuric reductase . FEBS Lett . 355 , 220 – 222 . Google Scholar Crossref Search ADS PubMed WorldCat [221] Rennex D. Cummings R.T. Pickett M. Walsh C.T. Bradley M. ( 1993 ) Role of tyrosine residues in Hg(II) detoxification by mercuric reductase from Bacillus sp. strain RC607 . Biochemistry 32 , 7475 – 7478 . Google Scholar Crossref Search ADS PubMed WorldCat [222] Moore M.J. Distefano M.D. Walsh C.T. Schiering N. Pai E.F. ( 1989 ) Purification, crystallization, and preliminary x-ray diffraction studies of the flavoenzyme mercuric ion reductase from Bacillus sp. strain RC607 . J. Biol. Chem . 264 , 14386 – 14388 . Google Scholar PubMed OpenURL Placeholder Text WorldCat [223] Miller, S.M. (1999) Construction of separate expression vectors for the catalytic core and the N-terminal metal binding domain of Tn501 mercuric ion reductase. In: Flavins and Flavoproteins 1999 (Ghisla, S., Kroneck, P., Macheroux, P. and Sund, H., Eds.), pp. 863–870. Rudolf Weber Agency for Scientific Publications, Berlin. [224] Miller S.M. Massey V. Williams C.H. Jr. Ballou D.P. Walsh C.T. ( 1991 ) Communication between the active sites in dimeric mercuric ion reductase: an alternating sites hypothesis for catalysis . Biochemistry 30 , 2600 – 2612 . Google Scholar Crossref Search ADS PubMed WorldCat [225] Misra T.K. Brown N.L. Fritzinger D.C. Pridmore R.D. Barnes W.M. Haberstroh L. Silver S. ( 1984 ) Mercuric ion-resistance operons of plasmid R100 and transposon Tn501: the beginning of the operon including the regulatory region and the first two structural genes . Proc. Natl. Acad. Sci. USA 81 , 5975 – 5979 . Google Scholar Crossref Search ADS WorldCat [226] Arnesano F. Banci L. Bertini I. Huffman D.L. O'Halloran T.V. ( 2001 ) Solution structure of the Cu(I) and apo forms of the yeast metallochaperone, Atx1 . Biochemistry 40 , 1528 – 1539 . Google Scholar Crossref Search ADS PubMed WorldCat [227] Banci L. Bertini I. Ciofi-Baffoni S. Huffman D.L. O'Halloran T.V. ( 2001 ) Solution structure of the yeast copper transporter domain Ccc2a in the apo and Cu(I)-loaded states . J. Biol. Chem . 276 , 8415 – 8426 . Google Scholar Crossref Search ADS PubMed WorldCat [228] Banci L. Bertini I. del Conte R. Markey J. Ruiz-Duenas F.J. ( 2001 ) Copper trafficking: the solution structure of Bacillus subtilis CopZ . Biochemistry 40 , 15660 – 15668 . Google Scholar Crossref Search ADS PubMed WorldCat [229] Gitschier J. Moffat B. Reilly D. Wood W.I. Fairbrother W.J. ( 1998 ) Solution structure of the fourth metal-binding domain from the Menkes copper-transporting ATPase . Nat. Struct. Biol . 5 , 47 – 54 . Google Scholar Crossref Search ADS PubMed WorldCat [230] Qian H. Sahlman L. Eriksson P.O. Hambraeus C. Edlund U. Sethson I. ( 1998 ) NMR solution structure of the oxidized form of MerP, a mercuric ion binding protein involved in bacterial mercuric ion resistance . Biochemistry 37 , 9316 – 9322 . Google Scholar Crossref Search ADS PubMed WorldCat [231] Rosenzweig A.C. Huffman D.L. Hou M.Y. Wernimont A.K. Pufahl R.A. O'Halloran T.V. ( 1999 ) Crystal structure of the Atx1 metallochaperone protein at 1.02 Å resolution . Struct. Fold. Des . 7 , 605 – 617 . Google Scholar Crossref Search ADS WorldCat [232] Hubbard T.J. Ailey B. Brenner S.E. Murzin A.G. Chothia C. ( 1999 ) SCOP: a Structural Classification of Proteins database . Nucleic Acids Res . 27 , 254 – 256 . Google Scholar Crossref Search ADS PubMed WorldCat [233] Newton, G.L. and Fahey, R.C. (1989) Glutathione in procaryotes. In: Glutathione: Metabolism and Physiological Functions (Viña, J., Ed.), pp. 69–77. CRC Press, Boca Raton, FL. [234] Newton G.L. Arnold K. Price M.S. Sherrill C. Delcardayre S.B. Aharonowitz Y. Cohen G. Davies J. Fahey R.C. Davis C. ( 1996 ) Distribution of thiols in microorganisms: mycothiol is a major thiol in most actinomycetes . J. Bacteriol . 178 , 1990 – 1995 . Google Scholar Crossref Search ADS PubMed WorldCat [235] Furukawa K. Tonomura K. ( 1971 ) Enzyme system involved in the decomposition of phenyl mercuric acetate by mercury-resistant Pseudomonas . Agric. Biol. Chem . 35 , 604 – 610 . OpenURL Placeholder Text WorldCat [236] Tezuka T. Tonomura K. ( 1978 ) Purification and properties of a second enzyme catalyzing the splitting of carbon-mercury linkages from mercury-resistant Pseudomonas K-62 . J. Bacteriol . 135 , 138 – 143 . Google Scholar PubMed OpenURL Placeholder Text WorldCat [237] Weiss A.A. Murphy S.D. Silver S. ( 1977 ) Mercury and organomercurial resistances determined by plasmids in Staphylococcus aureus . J. Bacteriol . 132 , 197 – 208 . Google Scholar PubMed OpenURL Placeholder Text WorldCat [238] Begley T.P. Walts A.E. Walsh C.T. ( 1986 ) Bacterial organo-mercurial lyase: overproduction, isolation and characterization . Biochemistry 25 , 7186 – 7192 . Google Scholar Crossref Search ADS PubMed WorldCat [239] Begley T.P. Walts A.E. Walsh C.T. ( 1986 ) Mechanistic studies of a protonolytic organomercurial cleaving enzyme: bacterial organomercurial lyase . Biochemistry 25 , 7192 – 7200 . Google Scholar Crossref Search ADS PubMed WorldCat [240] Gopinath E. Bruice T.C. ( 1987 ) Assistance of protodemercuration by bis-thiol ligation and nucleophilic catalysis. A model study which relates to the organomercurial lyase . J. Am. Chem. Soc . 109 , 7903 – 7905 . Google Scholar Crossref Search ADS WorldCat [241] Pitts K.E. Summers A.O. ( 2002 ) The roles of thiols in the bacterial organomercurial lyase (MerB) . Biochemistry 41 , 10287 – 10296 . Google Scholar Crossref Search ADS PubMed WorldCat [242] Ogawa H.I. Tolle C.L. Summers A.O. ( 1984 ) Physical and genetic map of the organomercury resistance (Omr) and inorganic mercury (Hgr) loci of the IncM plasmid R831b . Gene 32 , 311 – 320 . Google Scholar Crossref Search ADS PubMed WorldCat [243] Griffin H.G. Foster T.J. Silver S. Misra T.K. ( 1987 ) Cloning and DNA sequence of the mercuric and organomercurial-resistance determinants of plasmid pDU1358 . Proc. Natl. Acad. Sci. USA 84 , 3112 – 3116 . Google Scholar Crossref Search ADS WorldCat [244] Nucifora G. Chu L. Silver S. Misra T. ( 1989 ) Mercury operon regulation by the merR gene of the organomercurial resistance system of plasmid pDU1358 . J. Bacteriol . 171 , 4241 – 4247 . Google Scholar Crossref Search ADS PubMed WorldCat [245] Kholodii G. Gorlenko Z. Mindlin S. Hobman J. Nikiforov V. ( 2002 ) Tn5041-like transposons: molecular diversity, evolutionary relationships and distribution of distinct variants in environmental bacteria . Microbiology 148 , 3569 – 3582 . Google Scholar Crossref Search ADS PubMed WorldCat [246] Laddaga R.A. Chu L. Misra T.K. Silver S. ( 1987 ) Nucleotide sequence and expression of the mercurial-resistance operon from Staphylococcus aureus plasmid pI258 . Proc. Natl. Acad. Sci. USA 84 , 5106 – 5110 . Google Scholar Crossref Search ADS WorldCat [247] Wang Y. Moore M. Levinson H.S. Silver S. Walsh C. Mahler I. ( 1989 ) Nucleotide sequence of a chromosomal mercury resistance determinant from a Bacillus sp. with broad-spectrum mercury resistance . J. Bacteriol . 171 , 83 – 92 . TI - Bacterial mercury resistance from atoms to ecosystems JF - FEMS Microbiology Reviews DO - 10.1016/S0168-6445(03)00046-9 DA - 2003-06-01 UR - https://www.deepdyve.com/lp/oxford-university-press/bacterial-mercury-resistance-from-atoms-to-ecosystems-Sq2UcNuMa0 SP - 355 EP - 384 VL - 27 IS - 2-3 DP - DeepDyve ER -