RESEARCH/REVIEW ARTICLE

Some like it cold: microbial transformations of mercury in polar regions

Tamar Barkay,1,2 Niels Kroer2 & Alexandre J. Poulain3

1Department of Biochemistry and Microbiology, Rutgers University, 76 Lipman Dr, New Brunswick, NJ 08901, USA
2Department of Environmental Chemistry and Microbiology, National Environmental Research Institute, Aarhus University, Frederiksborgvej 399, Roskilde, DK-4000, Denmark
3Department of Biology, University of Ottawa, 30-268 Marie Curie, Ottawa, K1N 6N5, Ontario, Canada

Abstract

The contamination of polar regions with mercury that is transported from lower latitudes as inorganic mercury has resulted in the accumulation of methylmercury (MeHg) in food chains, risking the health of humans and wildlife. While production of MeHg has been documented in polar marine and terrestrial environments, little is known about the responsible transformations and transport pathways and the processes that control them. We posit that as in temperate environments, microbial transformations play a key role in mercury geochemical cycling in polar regions by: (1) methylating mercury by one of four proposed pathways, some not previously described; (2) degrading MeHg by activities of mercury resistant and other bacteria; and (3) carrying out redox transformations that control the supply of the mercuric ion, the substrate of methylation reactions. Recent analyses have identified a high potential for mercury-resistant microbes that express the enzyme mercuric reductase to affect the production of gaseous elemental mercury when and where daylight is limited. The integration of microbially mediated processes in the paradigms that describe mercury geochemical cycling is therefore of high priority especially in light of concerns regarding the effect of global warming and permafrost thawing on input of MeHg to polar regions.

Keywords
Microbiology; mercury biogeochemistry; redox transformations; polar regions; methylation

Correspondence
Tamar Barkay, Department of Biochemistry and Microbiology, Rutgers University, 76 Lipman Dr., New Brunswick, NJ 08901, USA. E-mail: barkay@aesop.rutgers.edu

(Published: 28 December 2011)

Citation: Polar Research 2011, 30, 15469, DOI: 10.3402/polar.v30i0.15469

Polar Research 2011. © 2011 T. Barkay et al. This is an open access article distributed under the terms of the Creative Commons Attribution-Noncommercial 3.0 Unported License (http://creativecommons.org/licenses/by-nc/3.0/), permitting all non-commercial use, distribution, and reproduction in any medium, provided the original work is properly cited.

Over the last few decades, concerns for the vulnerability of polar regions to organic and inorganic contaminants that originate in lower latitudes have increased. Mercury (Hg) is among the most serious of these contaminants due to its accumulation in polar food chains and the resulting health risks to both humans and wildlife (Macdonald et al. 2005; Dietz et al. 2009). Following natural and anthropogenic emissions, Hg is transported over long distances and globally distributed in its elemental form, Hg(0), which is also referred to as gaseous elemental Hg or GEM (Steffen et al. 2007; Pirrone et al. 2010). It is oxidized in the atmosphere and deposited via dry (aerosols) or wet (rain and snow) deposition to terrestrial and aquatic ecosystems. Asia is the dominant source of GEM to the Arctic (Durnford et al. 2010), rendering this region particularly vulnerable as emissions from Asia are expected to increase in coming decades (Streets et al. 2009). The Antarctic is contaminated from sources in Africa, Australia, and South America (Dommergue et al. 2010). Mercury deposition to polar regions is enhanced by springtime atmospheric Hg depletion events (MDE) in the High Arctic (Schroeder et al. 1998; Lindberg et al. 2002), sub-Arctic (Dommergue et al. 2003), and Antarctic (Ebinghaus et al. 2002) regions, resulting in rapid and massive deposition of ionic Hg, Hg(II), from the atmosphere (Brooks et al. 2006; Skov et al. 2006). This springtime deposition is thought to be due to the oxidation of GEM by halogen radicals and oxidized forms of halogens formed in sea salt aerosols by photochemical transformations (Lindberg et al. 2002; Brooks et al. 2006; Ariya et al. 2008).

How atmospherically deposited Hg(II) is converted to the potent neurotoxic compound methylmercury (MeHg) is the topic of this review. Our concerns in relation to MeHg production and its availability to polar food chains (Wren 1986) are due to human consumption of contaminated seals and whales (Macdonald et al. 2005) and to possible neurological damage in apex predators such as polar bears (Basu et al. 2009). In humans, MeHg manifests its toxicity in a variety of symptoms ranging from mild numbness of the extremities, blindness, impaired development of language, attention and memory skills (Krummel et al. 2005), and in severe cases, death (Clarkson 2002; Mergler et al. 2007).

Recent research has shown that Hg found in the highest trophic levels of Arctic food chains is almost exclusively present in the methylated form (Campbell et al. 2005; Loseto et al. 2008) and that blood and fatty tissues of native human populations have elevated levels of Hg (Van Oostdam et al. 2005; Butler Walker et al. 2006; Johansen et al. 2007; Donaldson et al. 2010). Thus, the impact of Hg contamination in the Arctic is similar to that described in temperate zones of the world, raising the critical question of how Hg(II), entering polar regions through atmospheric deposition, becomes available for accumulation as MeHg in food webs. The answers to this question are found in the dynamics of the polar Hg biogeochemical cycle (Fig. 1), i.e., within-ecosystem transformations play a critical role in the toxicity and distribution of Hg. Post-depositional Hg processes must therefore be understood before we can link Hg deposition to Hg burdens in polar biota (Macdonald & Loseto 2010).


Fig 1

Fig. 1  The biogeochemical cycle of mercury in coastal marine environments in polar regions. Major reaction and transport pathways, provided as numbers in parentheses in the figure, are: (1) atmospheric oxidation of Hg(0) to Hg(II); (2) photoreduction of newly deposited Hg(II) to Hg(0); (3) biological reduction of Hg(II) to Hg(0); (4) evasion of Hg(0) to the atmosphere; (5) methylation of Hg(II) to CH3Hg by sulfate-reducing bacteria (SRB) and iron-reducing bacteria (FeRB); (6) methylation of Hg(II) to CH3Hg by aerobic pathway and/or by photomethylation in snow; (7) methylation of Hg(II) to CH3Hg by algae and phytoplankton in the water column; (8) photochemical degradation of diMeHg in the atmosphere; (9) biological demethylation of CH3Hg to Hg(II); and (10) photochemical demethylation of CH3Hg in snow. Methylation pathways are highlighted by bold lettering. Note that these reactions and pathways may take place in various compartments of polar regions; for the sake of simplicity they are only marked in a representative compartment in the figure (see text for details). Dimethylsulfoniopropionate is abbreviated to DMSP, dimethylsulfide to DMS and methylsulfonic acid to MSA.

In temperate zones, microbial activities critically impact MeHg accumulation by carrying out biochemical transformations. Recent reviews on Hg cycling in the environment (Fitzgerald et al. 2007; Poissant et al. 2008; Selin 2009) and on the role of microbes (Barkay et al. 2003; Barkay et al. 2005) are available. Microbes are broadly distributed in polar environments, including air (Polunin & Kelly 1952), snow (Larose, Berger et al. 2010), coastal lagoons (Poulain, Ni Chadhain et al. 2007), soil (Connell et al. 2008), sea ice (Collins et al. 2010; Koh et al. 2010), marine sediments (Yergeau et al. 2009) and the water column (Galand et al. 2009). Bacteria and bacteriophages have also been documented in frost flowers (Bowman & Deming 2010), where Hg concentrations (as high as 5 nmol L−1 or 1 µg L−1) are more than 10 times higher than in MDE snow (Douglas et al. 2005; Douglas et al. 2008) and almost a 1000-fold higher than in Arctic inland locations (St. Louis et al. 2005). Because microbial activities have been documented in samples collected in both the Arctic (Kirchman et al. 2007; Yergeau et al. 2009) and the Antarctic (Manganelli et al. 2009), albeit at rates lower than those in temperate regions, research on microbial activities in polar regions should be an important component of efforts directed towards the understanding of how Hg biogeochemistry is related to MeHg accumulation. This need is highlighted by the paucity of published peer-reviewed publications on the interactions of microorganisms with Hg in polar regions. While the numbers of papers describing Hg or microbes in polar regions are in the hundreds, search engines have only picked up a single publication when the terms “microbes” OR “bacteria” OR “archaea” AND “mercury” AND “arctic” were used, and none when “antarctic” was replaced with “arctic” (Fig. 2).


Fig 2

Fig. 2  The number of papers retrieved on 11 October 2010 from the ISI Web of Knowledge database, using the keywords indicated. The search was performed using Boolean operators to avoid references to unrelated topics. The descriptor “microbes” is used for clarity and is based on a search that was performed using the query Microbes OR Bacteria OR Archaea AND all other terms as indicated in the figure.

Here we update our 2007 review paper and consider the most recent information on Hg in cold environments together with relevant information from research on Hg and microbiology in temperate environments. We synthesize these sources of information to propose junctures where microbes critically affect the geochemical cycle of Hg in polar regions (Fig. 1) and identify research questions that address gaps in our understanding of how microbes modulate the toxicity and mobility of Hg in the Arctic and Antarctic regions (Table 1).


Table 1 A summary of the uniqueness of microbial transformations and research questions whose answers would enhance our understanding of Hg biogeochemistry in polar regions.

Microbial transformation What is unique about this transformation in polar regions Questions/research needs
Methylation
  • Presence of diMeHg in coastal water (Pongratz & Heumann 1999; Kirk et al. 2008)

  • Snow as a matrix for Hg transformations and MeHg transport (Constant et al. 2007).

  • Marine sources for MeHg deposition in coastal regions (St. Louis et al. 2005; Larose, Dommergue et al. 2010)

  • What are the pathways for methylation and what fraction of the deposited Hg is being methylated?

  • What are the pathways for aerobic Hg methylation?

  • Who methylates Hg in polar regions?

  • What is the effect of permafrost thawing on methylation rate and subsequent input of MeHg to polar regions?

Demethylation
  • Oxidation of C1 compounds is slow in high latitudes (Hines & Duddleston 2001)

  • mer gene expression in Arctic biomass (Poulain, Ni Chadhain et al. 2007)

  • Photoreduction of MeHg in epilimnetic lake water (Hammerschmidt & Fitzgerald 2006)

  • What are the pathways for the degradation of MeHg in polar regions?

Hg(II) reduction
  • Accumulation of dissolved gaseous Hg under sea ice (Andersson et al. 2008)

  • Hg-resistant bacteria are common in snowpacks (Møller et al. 2011)

  • High bioavailability of Hg(II) in freshly deposited snow (Lindberg et al. 2002), Barkay & Kroer (unpubl. data)

  • Interactions of microbes with Hg in structured environments

  • mer gene expression in Arctic biomass (Poulain, Ni Chadhain et al. 2007) and its impact on Hg(II) reduction

  • Development of psychrophilic Hg biosensors

  • The interactions of microbes in sea ice with Hg; role of exopolysaccharide production

  • Measurement of Hg concentrations in the complex sea ice matrix

  • Further assess the evolution of Hg resistance in polar areas

Hg(0) oxidation
  • High chloride concentrations in coastal marine environments induce abiotic oxidation of Hg(0)

  • A better understanding of Hg(0) oxidation in Hg biogeochemistry


Microbial transformations of mercury in polar environments

Our current view of the role of microorganisms in the cycling of Hg in the environment is based on studies that were initiated by the discovery of the toxicity of MeHg to consumers of contaminated fish and shellfish in the 1960s (Westöö 1966). Results from environmental, geochemical, microbiological, biochemical, and molecular studies have converged to establish our current view of the Hg biogeochemical cycle (Barkay et al. 2005; Fitzgerald et al. 2007; Selin 2009). Within that paradigm, microbes impact the production of MeHg directly by methylation and demethylation processes, and indirectly by controlling the supply of Hg(II), the substrate for methylation, by carrying out redox transformations that affect transitions between Hg(II), and Hg(0). These transformations and how they are likely to be impacted by the unique conditions of cold environments are discussed below.

Hg(II) methylation

Anaerobic microbes have been known for over 40 years to methylate Hg (Jensen & Jernelöv 1969) and for the last 25 years this activity has been attributed to sulfate-reducing bacteria (SRB) in anoxic environments (Compeau & Bartha 1985; Gilmour et al. 1992; King et al. 2000). The mechanism of methylation may (Choi et al. 1994) or may not (Ekstrom et al. 2003) be related to the production of acetyl coenzyme A and methylcobalamine (Ekstrom & Morel 2008). More recently, methylation by some iron-reducing bacteria (FeRB) has been suggested (Fleming et al. 2006; Kerin et al. 2006), although when tested under environmentally relevant conditions, only SRB produced significant amounts of MeHg (Ranchou-Peyruse et al. 2009). Methylation of Hg(II) by abiotic processes (Weber 1993; Siciliano et al. 2005) may be indirectly related to biological activities because of its dependence on biological products such as dissolved organic matter.

Formation of MeHg in the Arctic has been documented in wetland soils (Loseto, Siciliano et al. 2004; Oiffer & Siciliano 2009) and streams (Loseto, Lean et al. 2004), in snow (Constant et al. 2007), in freshwater ponds (St. Louis et al. 2005), in the marine water column (Kirk et al. 2008), and in lakes and tundra watersheds (Hammerschmidt et al. 2006). Based on several considerations we suggest that at least four different methylation pathways contribute to MeHg formation and accumulation is polar regions. These considerations include: (1) the distribution of MeHg and microbial communities in various compartments of the cryosphere; (2) the unique physical properties of polar environments; (3) advances in elucidating the microbial cold way of life using genomic approaches (Methe et al. 2005); and (4) knowledge of the biochemistry and physiology of microbial transmethylation reactions. With the exception of methylation by SRB and FeRB, evidence for the proposed pathways is lacking. They are highlighted here because their occurrence in polar environments is plausible when available data from polar regions are synthesized with our current understanding of the chemistry and biochemistry of Hg methylation.

Methylation by SRB and FeRB

The large component of coastal shelves in the Arctic Ocean (Macdonald & Loseto 2010) and the high summer productivity of coastal lagoons (Galand et al. 2008) highlight the likely importance of methylation by SRB and FeRB in anoxic sediments as a possible source of MeHg (pathway 5 in (Fig. 1). However, Loseto, Siciliano et al. (2004), who detected low abundance of SRB and failed to detect Deltaproteobacteria and genes encoding for the disulfite reductase enzyme in soil DNA extracts, concluded that methylation was not mediated by SRB. This conclusion may have been premature because if methylating SRB are a minor component in the soil community, the sensitivity of the molecular methods may have not been sufficient to detect them. For example, we were recently able to attribute methylation in an Adirondack wetland to SRB only when experiments with metabolic enhancers and inhibitors and highly sensitive molecular methods were employed (Yu et al. 2010). Therefore, the involvement of SRB in methylation in polar regions, especially in anoxic sediments of coastal environments, where sulfate reduction is likely the dominant respiratory pathway, remains to be examined. This involvement is supported by observations that SRB are abundant in Arctic coastal marine sediments such as in Svalbard, Norway (Ravenschlag et al. 2001), and in Antarctic sediments (Purdy et al. 2003) and that psychrophilic SRB isolated from the same sediments actively reduced sulfate at in situ temperatures (Knoblauch et al. 1999; Bruchert et al. 2001). To the best of our knowledge, the role of FeRB in methylation in polar regions has not been explored though iron, like sulfate, reduction readily occurs in cold environments (Finke et al. 2007).

Methylation in the marine water column

The production of mono- and dimethylmercury (diMeHg) in the Arctic marine environment (pathway 7 in (Fig. 1), recently documented by Kirk et al. (2008) in mid- to bottom depth in the Canadian Arctic Archipelago and in the Hudson Strait and Hudson Bay, is likely a part of the larger story of MeHg production in the marine water column thought to be associated with the remineralization of particulate organic carbon in oxygen minima zones (Monperrus et al. 2007; Cossa et al. 2009; Sunderland et al. 2009). Which organisms are involved in marine water column methylation is currently not known, but these may not be anaerobic microbes as suggested by the failure to detect such microbes at depth where MeHg accumulated (Malcolm et al. 2010). Methylation by phytoplankton and/or their exudates is a possibility as previously reported in a coastal Antarctic water column (Pongratz & Heumann 1999). A possible mechanism for the phytoplankton-associated methylation was very recently proposed by Larose and co-workers (Larose, Dommergue et al. 2010) implicating transmethylation reactions that are involved in the degradation of the phytoplankton osmolyte dimethylsulfoniopropionate (DMSP) (Bentley & Chasteen 2004). Together, these studies challenge the current paradigm that only anaerobic conditions support significant MeHg build up (or net rates of methylation) and underscores the need for more discovery based fundamental research examining mechanistics aspects of Hg methylation.

Snowpacks: in-snow methylation vs. transport from marine sources

One unique aspect of MeHg accumulation in coastal Arctic environments is a high concentration of MeHg in meltwater at the initiation of snowmelt (Loseto, Lean et al. 2004; St. Louis et al. 2005) suggesting accumulation of MeHg in snowpacks where anaerobic environments are uncommon (pathway 6 in (Fig. 1). Positive correlations between MeHg and chloride or methanesulfonates, a product of DMSP degradation (Bentley & Chasteen 2004), and total Hg and chloride (St. Louis et al. 2007) in snowpacks suggest a marine source for Hg. We can speculate that MeHg and diMeHg produced in the marine water column (Kirk et al. 2008; Cossa et al. 2009) may evade from productive leads and polynyas followed by deposition onto sea ice and terrestrial systems (St. Louis et al. 2005). The photodegradation of the highly volatile diMeHg to MeHg in the atmosphere (Niki et al. 1983) could be a part of this process.

In-snow methylation of bioavailable Hg(II), however, cannot be ruled out. For example, methylation in tundra snowpacks was suggested by correlations between the proportion of total Hg as MeHg and heterotrophic bacterial counts and concentrations of suspended solids (Constant et al. 2007; Kirk et al. 2008). Experiments using bioreporters (Selifonova et al. 1993; Golding et al. 2002) suggested that a significant proportion of Hg(II) deposited during MDE in Barrow, Alaska, was bioavailable (Scott 2001; Lindberg et al. 2002). Similarly, five out of 12 surface/top layer snow samples collected during or following a snowstorm at Station Nord, north-east Greenland, in spring 2010, had significant amounts of bioavailable Hg (Barkay & Kroer unpubl. data). Moreover, organic compounds, such as dicarboxylic acids, are present in Arctic snow (Kawamura et al. 1996) and may serve as a carbon and energy source for microorganisms (Amato et al. 2007) that may be involved in methylation processes. Our direct bacterial counts showed 2×105 cells per ml of melted snow from the Canadian High Arctic and 1×103 cells per ml of melted snow from north-east Greenland (Møller et al. 2011) while melted snow from Antarctica's dry valleys had 200–5000 cells per ml (Alfreider et al. 1996; Carpenter et al. 2000; Segawa et al. 2005). Amato et al. (2007) reported 2×104 and 6×104 cells per ml in snow accumulated over a glacier on Spitsbergen, Svalbard, and in a seasonal snowpack bordering the Arctic Ocean, respectively. Microbes in snow may be metabolically active, as indicated by the reduction of INT, a respiratory indicator (Alfreider et al. 1996) and by low, but detectable, levels of protein and nucleic acid synthesis at in situ temperatures (Carpenter et al. 2000). This suggests the possibility that microbes in snow may methylate Hg. This proposition, like methylation in the oxygenated marine column (see above), implies methylation by aerobic microorganisms. Many aerobic microorganisms may methylate Hg, e.g., bacteria belonging to the Pseudomonas, Enterobacter, Bacillus, and Staphylococci genera and fungi such as Aspergillus niger, Scopulariopsis brevicattlis and Saccharomyces cerevisiae (Vonk & Sijpesteijn 1973), and the activity of these microbes may be environmentally relevant but remains to be demonstrated.

Photomethylation

It has long been known that MeHg may be formed in solutions containing various organic molecules in response to light (Hayashi et al. 1977) and more recently Siciliano et al. (2005) showed that photomethylation in northern temperate ecosystems depended on the presence and size of dissolved organic matter. This process may affect MeHg formation in snow (pathway 6 in (Fig. 1) and other cold environments where biological processes produce dissolved organic matter (Calace et al. 2005).

As has been the case with studies of methylation in temperate regions, direct experimentation using pure cultures of active microbes (Choi et al. 1994), laboratory incubations (Yu et al. 2010), and testing in intact and/or manipulated environmental incubations (Hammerschmidt et al. 2006; Monperrus et al. 2007) are needed to distinguish the relative importance of the four proposed methylation pathways to the accumulation of MeHg in polar regions. This research will benefit greatly from the availability of the sequenced genomes of psychrophilic microbes (Methe et al. 2005) and the metagenomes of microbial communities from cold environments (Larose, Berger et al. 2010). We hypothesize that methylation by anaerobic bacteria is prominent considering the large magnitude of coastal shelves and inputs from river discharge to the high Arctic (Macdonald & Loseto 2010). Yet, considering that both poles, the Arctic in particular, are highly influenced by processes in the marine environment, methylation in water column, and by aerobic microbes, may be a significant contributor to the MeHg pool in polar regions.

Methylmercury degradation

Because they consume MeHg, demethylation reactions impact net methylation rates and thus the net production of this neurotoxic substance. Three demethylation processes—photodegradation (Sellers et al. 1996) and two microbially mediated processes (Schaefer et al. 2004; Barkay et al. 2005)—are known. Photodegradation, a process mediated by ultraviolet radiation (Lehnherr & St. Louis 2009) and enhanced by the presence of organic ligands (Zhang & Hsu-Kim 2010), is the dominant mechanism for demethylation in surface water. It has been invoked as the sole process responsible for the degradation of MeHg in the eplimnitic water of a highly oligotrophic freshwater Arctic lake (Hammerschmidt & Fitzgerald 2006). To the best of our knowledge, MeHg degradation has not been examined in the euphotic zone of sediments or samples from coastal marine environments in polar regions.

Microbial pathways for the degradation of MeHg are distinguished by the redox state of the gaseous carbon products of demethylation. In reductive demethylation, methane is produced and in the oxidative process the product is both carbon dioxide and methane. We (Schaefer et al. 2004) and others (Marvin-Dipasquale et al. 2000; Gray et al. 2004) have shown that the choice between these processes is to a large extent controlled by environmental factors. Reductive demethylation is mediated by the organomercury lyase enzyme, which is a part of the Hg resistance (mer) system in bacteria (see below). This process is favoured at a high redox potential and high concentrations of Hg since expression of mer operon genes is induced by inorganic divalent Hg (Schaefer et al. 2004; Barkay et al. 2010). Oxidative demethylation is favored at low redox potentials and at a broad range of Hg concentrations and is most likely related to C1-pathways in anaerobic prokaryotes (Marvin-Dipasquale & Oremland 1998). The occurrence and rates of C1 metabolism in microbes from cold environments have been getting a lot of attention due to anticipated effects of global warming on the release of carbon from large frozen reservoirs in permafrost and polar tundra. While methanogenesis (Rivkina et al. 2004; Berestovskaya et al. 2005) and methanotrophy (Berestovskaya et al. 2005) were detected in permafrost, rates were drastically impacted by a drop in the incubation temperature. Moreover, degradation of C1 compounds such as methylbromide or acetate, common in temperate soils (Hines et al. 1998), is rarely observed at high latitudes proximal to polar areas (Hines & Duddleston 2001). Based on these observations the likelihood for oxidative MeHg degradation in polar regions is currently low but may increase should the carbon cycle be altered by warmer conditions. Nevertheless, demethylation plays an important role in determining MeHg production and availability to food chains and its occurrence and mechanisms in cold environments need to be addressed.

Redox transformations of inorganic Hg

Redox transformations between the ionic and elemental Hg forms affect MeHg production by controlling the amount of the substrate that is available for methylation (Fitzgerald et al. 1991). Among the reduction processes, photoreduction dominates in surface water (Krabbenhoft et al. 1998; Amyot et al. 2004; O'Driscoll et al. 2004; Poulain, Amyot et al. 2004; Garcia, Amyot et al. 2005; Zhang et al. 2006) and snow (Lalonde et al. 2002; Lalonde et al. 2003) and is thought to result in the evasion of most of the Hg that is deposited onto snow (Lindberg et al. 2002; Dommergue et al. 2003) or condensed into frost flowers (Douglas et al. 2008). The impact of photoreduction on Hg that is deposited during springtime MDE was recently confirmed by showing a large negative mass-independent fractionation of Hg isotopes (Sherman et al. 2010) thought to be exclusively induced by light mediated reactions (Bergquist & Blum 2007; Kritee et al. 2009).

There are several microbial Hg reduction processes and chief among them is mediated by the inducible Hg resistance (mer) operon in Hg-resistant bacteria, which impacts the partition of Hg into the gaseous phase in some environments (Barkay 1987; Barkay et al. 2005). The enzyme mercuric reductase, MerA, encoded by the merA gene, is the core function of the mer operon, which also encodes for Hg transport and for an elaborate system that regulates expression of the operon. Some operons also encode for the organomercury lyase and microbes carrying such mer operons reductively degrade MeHg (see above). The mer operon is broadly distributed among Bacteria (Barkay et al. 2010) and Archaea (Simbahan et al. 2005) from diverse environments (Osborn et al. 1997; Barkay et al. 2010). The presence of Hg resistant bacteria in samples from polar environments (Møller et al. 2011) and the demonstration of merA gene expression in samples from the High Arctic (Poulain, Ni Chadhain et al. 2007) suggest that Hg resistant microbes may be endemic and active in cold regions, a conclusion that is also supported by the presence of mer gene homologs in the genomes of several psychrophilic bacteria from polar environments (Barkay et al. 2010). Resistant bacteria accounted for 0–31% of the total number of the cultured bacteria in High Arctic snow and for approximately 2% in sea ice brine and freshwater (Møller et al. 2011) and for up to 3% in permafrost sediments (Petrova et al. 2002; Mindlin et al. 2005), and for 1–68% in Antarctic seawater (De Souza et al. 2006; Miller et al. 2009).

The regulator of mer expression, MerR, plays a critical role in determining where and under which conditions Hg(II) reduction by MerA occurs. Expression of the mer-operon is repressed in the absence of Hg and is quantitatively induced in its presence (Summers 1992; Brown et al. 2003). Because of this requirement for induction, MerA-mediated reduction has been considered of little relevance to transformations of Hg in natural environments (Morel et al. 1998). Indeed, a series of studies, performed in several environments that were impacted by various sources of Hg, showed mRNA transcripts of the merA gene in highly contaminated environments, whereas microbial biomass from environments with low levels of contamination contained low to non-detectable levels of these transcripts (Nazaret et al. 1994; Hines et al. 2000; Poulain, Amyot et al. 2004; Schaefer et al. 2004). Based on these observations one would not expect merA expression and reduction of Hg(II) by polar microbial communities where Hg concentrations are in the pM range during most times of the year (Steffen et al. 2002; St. Louis et al. 2005). We were, therefore, surprised when merA transcripts were detected in microbial biomass associated with algae that were collected in a coastal lagoon and a sea-ice lead in the Canadian High Arctic in the summer of 2005 where total Hg concentrations were ca. 10 pM (Poulain, Ni Chadhain et al. 2007). Numerous reasons may account for this apparent discrepancy. The absolute Hg concentrations required for mer induction depend on the complicated issue of bioavailability and how it is impacted by interactions with ligands in the environment ( 1997; Crespo-Medina et al. 2009). In environments with low concentrations of ligands, induction may take place at very low Hg concentrations. For example, induction of the mer-lux bioreporter in laboratory incubations was documented at sub-pM Hg concentrations when “clean conditions” were employed (Kelly et al. 2003). The slow rates of transcript degradation in cold environments (Vlassov et al. 2005), might furthermore explain the detection of merA transcripts in polar microbiota.

Alternatively, the highly heterogeneous nature of microbial habitats in polar regions may lead to locally high concentrations of Hg in micro-niches where mer induction may take place. The effect of heterogeneous micro-environments on the distribution of Hg and on selection of resistant bacteria was recently demonstrated (Slater et al. 2008); selection extended to a distance of <500 µm from Hg foci created by the impregnation of fiber with mercuric chloride (Slater et al. 2010). Sea-ice, the habitat for most of the microbial biomass in coastal polar environments, may contain niches where both Hg and microorganisms are concentrated. It is likely that Hg, like other solutes in sea-ice (Eicken 2003), is highly concentrated in brine channels where actively metabolizing microorganisms were documented (Deming 2002; Junge et al. 2004). Our hypothesis on the localized proximity of microbes to Hg in brine channels is also supported by the observations that microbes in brine channels during winter are associated with particles (Junge et al. 2004), that a significant fraction of atmospherically derived Hg is bound to particles (Schroeder & Munthe 1998), and that Hg in snow—especially in marine environments—is almost exclusively associated with particles (Poulain, Garcia et al. 2007). However, the discovery of copious production of exopolysaccharides by microbes in sea-ice (Krembs & Deming 2008), proposed as a cryoprotection mechanism (Marx et al. 2009), may suggest an alternative mechanism for Hg tolerance whereby Hg is sequestered extracellularly as has been shown for other metals in other environments (Teitzel & Parsek 2003). The possibility that resistance to Hg among sea-ice bacteria in brine channels is not mediated by mer systems is supported by a low number of Hg resistant culturable bacterial counts in brine samples extracted from sea-ice at Station Nord in north-east Greenland (Møller et al. 2011).

Bioreduction of Hg, unrelated to the mer system, may be associated with the activity of microorganisms in fresh and salt waters via pathways still to be determined. These could be related to both heterotrophic and/or phototrophic activities (Ben-Bassat & Mayer 1978; Mason et al. 1995; Poulain, Amyot et al. 2004; Rolfhus & Fitzgerald 2004; Wiatrowski et al. 2006; Wiatrowski et al. 2009).

How significant is microbial reduction of Hg(II) in Hg geochemistry in polar regions? Numbers of merA transcripts in Arctic microbial biomass (Poulain, Ni Chadhain et al. 2007) and numbers of Hg resistant bacteria in snowpacks (Møller et al. 2011) were used to answer this question. Using Acuchem modeling software (Braun et al. 1988) and a custom-designed kinetic code, Poulain, Ni Chadhain et al. (2007) showed that at equilibrium and when 5% of bacterial cells were considered active 65% of the elemental Hg (Hg[0]) was biogenic at the surface of the Arctic ocean while at a depth of 10 m with diminishing UVA and UVB radiation this fraction increased to 94%. Likewise, an almost 20-fold increase in the potential reduction rate was predicted in snowpacks at Station Nord with sampling depth increasing from about 83 to 105 cm. Comparison with reduction rates measured in snow from the Canadian High Arctic (Dommergue et al. 2003) suggested that an average of up to 2% of the total reduction could be biological and that bacterial reduction became increasingly important with snow depth (Møller et al. 2011). There is therefore a potential for microbial reduction to affect Hg mobility in the Arctic, especially at depth and under sea ice where light and the flux of dissolved gaseous Hg (DGM) to the atmosphere are limited. This conclusion is consistent with observations of enhanced DGM concentrations recorded underneath sea ice (Andersson et al. 2008). Our results and analyses suggest that most of the DGM pool in the Arctic Ocean could be of a microbial origin. Further studies should expend these preliminary findings.

The microbial oxidation of Hg(0) to Hg(II) is the part of the Hg biogeochemical cycle about which we know the least. To date, most research efforts have examined abiotic mechanisms of light and dark oxidation (Lalonde et al. 2001; Lalonde et al. 2004; Poulain, Lalonde et al. 2004; Raofie & Ariya 2004; Sheu & Mason 2004; Garcia, Poulain et al. 2005; Whalin & Mason 2006). Bacterial enzymes known for their role in the response to oxidative damage, such as catalases and hydroperoxidases, oxidize Hg(0) in organisms that are common in natural waters and soils (Smith et al. 1998). Furthermore, Siciliano et al. (2002) related specific rates of Hg(0) oxidation by lake microbial biomass to variations in DGM concentrations. How these microbially-mediated oxidative processes affect Hg speciation in polar regions, and especially their impact on the fate of DGM, has not been examined.

Conclusions and future needs

The study of Hg (micro)biogeochemistry in polar environments is at its early stages, but the synthesis of information available from temperate regions together with what we know about the distribution of Hg in polar regions and about microbiology in cold environments points to the uniqueness of Hg cycling in polar regions (Table 1). As in temperate environments, MeHg is accumulated by aquatic food chains but the methylation pathways themselves and the sites where methylation occurs may differ from those in lower latitudes. A particularity of polar ecosystems is the enhanced vulnerability of marine and coastal environments to Hg accumulation due to enhanced deposition during springtime.

Global warming poses a major challenge to the management of Hg contamination in polar regions. Increased temperatures are likely to directly affect Hg biogeochemistry by enhancing the rates of microbial transformations and yearly productivity as polar summers are lengthened. In addition, open waters created with the accelerated melting of sea ice are likely to result in higher inputs of halogen aerosols to the atmosphere and the subsequent enhanced deposition of RGM with precipitation. The impact of these changes on both microbial and abiotic methylation as well as MeHg degradation and redox transformations of inorganic Hg will determine future trends in MeHg accumulation in polar regions.

Thawing permafrost may be an increasing source of MeHg to polar ecosystems. One may expect an increased production of MeHg in polar regions as a consequence of global warming considering the known relationship of enhanced methylation with increased oxidation of organic matter (Kelly et al. 1997; St. Louis et al. 2004) and the increased cycling of carbon (Davidson & Janssens 2006; Heimann & Reichstein 2008) together with the release of Hg from peat (Klaminder et al. 2008) when permafrost thaws. Considering the enormous magnitude of carbon that is sequestered in permafrost and the projection for rapid permafrost thawing (Lawrence & Slater 2005), an evaluation of how this change can affect Hg biogeochemistry is needed.

A better description and understanding of Hg transport and transformations in sea-ice microbial habitats is warranted. These marine environments are characterized by spatially and temporally fractured unique environments in terms of their physical, chemical, and biological features. These niches may alter, or modulate, the pathways of microbial transformations of Hg relative to their characteristics in temperate environments. Our current state of knowledge provides us with a starting point for studies on Hg transformations in polar regions, and such studies promise to add new dimensions to our perception of the mechanisms and pathways that determine Hg toxicity and facilitate life in its presence.

Acknowledgements

Henrik Skov is thanked for valuable comments on atmospheric Hg transformations and transport. The authors’ research on Hg biogeochemistry is supported by the Environmental Remediation Science Program, Biological and Environmental Research, of the US Department of Energy to TB, the Natural Science and Engineering Research Council of Canada to AJP and a Marie Curie Actions—International Incoming Fellowship to TB and NK.

References

Alfreider A., Pernthaler J., Amann R., Sattler B., Glockner F.O., Wille A. & Psenner R. 1996. Community analysis of the bacterial assemblages in the winter cover and pelagic layers of a high mountain lake by in situ hybridization. Applied and Environmental Microbiology 62, 2138–2144.

Amato P., Hennebelle R., Magand O., Sancelme M., Delort A.M., Barbante C., Boutron C. & Ferrari C. 2007. Bacterial characterization of the snow cover at Spitzberg, Svalbard. FEMS Microbiology Ecology 59, 255–264. [Crossref]

Amyot M., Southworth G., Lindberg S.E., Hintelmann H., Lalonde J.D., Ogrinc N., Poulain A.J. & Sandilands K.A. 2004. Formation and evasion of dissolved gaseous mercury in large enclosures amended with (HgCl2)-Hg-200. Atmospheric Environment 38, 4279–4289. [Crossref]

Andersson M.E., Sommar J., Gardfeldt K. & Lindqvist O. 2008. Enhanced concentrations of dissolved gaseous mercury in the surface waters of the Arctic Ocean. Marine Chemistry 110, 190–194. [Crossref]

Ariya P.A., Skov H., Grage M.L. & Goodsite M.E. 2008. Gaseous elemental mercury in the ambient atmosphere: review of the application of theoretical calculations and experimental studies for determination of reaction coefficients and mechanisms with halogens and other reactants. Advances in Quantum Chemistry 55, 43–54. [Crossref]

Barkay T. 1987. Adaptation of aquatic microbial communities to Hg2+ stress. Applied and Environmental Microbiology 53, 2725–2732.

Barkay T., Gillman M. & Turner R.R. 1997. Effects of dissolved organic carbon and salinity on bioavailability of mercury. Applied and Environmental Microbiology 63, 4267–4271.

Barkay T., Kritee K., Boyd E. & Geesey G. 2010. A thermophilic bacterial origin and subsequent constraints by redox, light and salinity on the evolution of the microbial mercuric reductase. Environmental Microbiology 12, 2904–2917. [Crossref]

Barkay T., Miller S.M. & Summers A.O. 2003. Bacterial mercury resistance from atoms to ecosystems. FEMS Microbiology Reviews 27, 355–384. [Crossref]

Barkay T., Schaefer J.K., Poulain A.J. & Amyot M. 2005. Microbial transformations in the mercury geochemical cycle. Geochimica et Cosmochimica Acta 69, A702–A702.

Basu N., Scheuhammer A.M., Sonne C., Letcher R.J., Born E.W. & Dietz R. 2009. Is dietary mercury of neurotoxicological concern to wild polar bears (Ursus maritimus)? Environmental Toxicology and Chemistry 28, 133–140. [Crossref]

Ben-Bassat D. & Mayer A.M. 1978. Light induced Hg volatilization and O2 evolution in Chlorella and the effect of DCMU and methylamine. Physiologia Plantarum 42, 33–38. [Crossref]

Bentley R. & Chasteen T.G. 2004. Environmental VOSCs—formation and degradation of dimethyl sulfide, methanethiol and related materials. Chemosphere 55, 291–317. [Crossref]

Berestovskaya Y.Y., Rusanov I.I., Vasil'eva L.V. & Pimenov N.V. 2005. The processes of methane production and oxidation in the soils of the Russian Arctic tundra. Microbiology 74, 221–229. [Crossref]

Bergquist B.A. & Blum J.D. 2007. Mass-dependent and -independent fractionation of Hg isotopes by photoreduction in aquatic systems. Science 318, 417–420. [Crossref]

Bowman J.S. & Deming J.W. 2010. Elevated bacterial abundance and exopolymers in saline frost flowers and implications for atmospheric chemistry and microbial dispersal. Geophysical Research Letters 37, L13501, doi: 10.1029/2010GL043020. [Crossref]

Braun W., Herron J.T. & Kahaner D.K. 1988. Acuchem: a computer-program for modeling complex chemical-reaction systems. International Journal of Chemical Kinetics 20, 51–62. [Crossref]

Brooks S., Saiz-Lopez A., Skov H., Lindberg S., Plane J.M.C. & Goodsite M.E. 2006. The mass balance of mercury in the springtime Arctic environment. Geophysical Research Letters 33, L13812, doi: 10.1029/2005GL025525. [Crossref]

Brown N.L., Stoyanov J.V., Kidd S.P. & Hobman J.L. 2003. The MerR family of transcriptional regulators. FEMS Microbiology Reviews 27, 145–163. [Crossref]

Bruchert V., Knoblauch C. & Jørgensen B.B. 2001. Controls on stable sulfur isotope fractionation during bacterial sulfate reduction in Arctic sediments. Geochimica et Cosmochemica Acta 65, 763–776. [Crossref]

Butler Walker J., Houseman J., Seddon L., McMullen E., Tofflemire K., Mills C., Corriveau A., Weber J.P., LeBlanc A., Walker M., Donaldson S.G. & Van Oostdam J. 2006. Maternal and umbilical cord blood levels of mercury, lead, cadmium, and essential trace elements in Arctic Canada. Environmental Research 100, 295–318. [Crossref]

Calace N., Cantafora E., Mirante S., Petronio B.M. & Pietroletti M. 2005. Transport and modification of humic substances present in Antarctic snow and ancient ice. Jounral of Environmental Monitoring 7, 1320–1325. [Crossref]

Campbell L.M., Norstrom R.J., Hobson K.A., Muir D.C.G., Backus S. & Fisk A.T. 2005. Mercury and other trace elements in a pelagic Arctic marine food web (Northwater Polynya, Baffin Bay). Science of the Total Environment 351, 247–263. [Crossref]

Carpenter E.J., Lin S.J. & Capone D.G. 2000. Bacterial activity in South Pole snow. Applied and Environmental Microbiology 66, 4514–4517. [Crossref]

Choi S.C., Chase T. & Bartha R. 1994. Metabolic pathways leading to mercury methylation in Desulfovibrio desulfuricans LS. Applied and Environmental Microbiology 60, 4072–4077.

Clarkson T.W. 2002. The three modern faces of mercury. Environmental Health Perspectives 110(Supplement 1), 11–23. [Crossref]

Collins R.E., Rocap G. & Deming J.W. 2010. Persistence of bacterial and archaeal communities in sea ice through an Arctic winter. Environmental Microbiology 12, 1828–1841. [Crossref]

Compeau G.C. & Bartha R. 1985. Sulfate-reducing bacteria: principle methylators of mercury in anoxic estuarine sediment. Applied and Environmental Microbiology 50, 498–502.

Connell L., Redman R., Craig S., Scorzetti G., Iszard M. & Rodriguez R. 2008. Diversity of soil yeasts isolated from South Victoria Land, Antarctica. Microbial Ecology 56, 448–459. [Crossref]

Constant P., Poissant L., Villemur R., Yumvihoze E. & Lean D. 2007. Fate of inorganic mercury and methyl mercury within the snow cover in the Low Arctic tundra on the shore of Hudson Bay (Québec, Canada). Journal of Geophysical Research—Atmospheres 112, D08309 [Crossref]

Cossa D., Averty B. & Pirrone N. 2009. The origin of methylmercury in open Mediterranean waters. Limnology and Oceanography 54, 837–844. [Crossref]

Crespo-Medina M., Chatziefthimiou A.D., Bloom N.S., Luther G.W. III, Reinfelder J.R., Vetriani C. & Barkay T. 2009. Adaptation of chemosynthetic microorganisms to elevated mercury concentrations in deep-sea hydrothermal vents. Limnology & Oceanography 54, 41–49. [Crossref]

Davidson E.A. & Janssens I.A. 2006. Temperature sensitivity of soil carbon decomposition and feedbacks to climate change. Nature 440, 165–173. [Crossref]

De Souza M.J., Nair S., Bharathi P.A.L. & Chandramohan D. 2006. Metal and antibiotic-resistance in psychrotrophic bacteria from Antarctic marine waters. Ecotoxicology 15, 379–384. [Crossref]

Deming J.W. 2002. Psychrophiles and polar regions. Current Opinion in Microbiology 5, 301–309. [Crossref]

Dietz R., Outridge P.M. & Hobson K.A. 2009. Anthropogenic contributions to mercury levels in present-day Arctic animals—a review. Science of the Total Environment 407, 6120–6131. [Crossref]

Dommergue A., Ferrari C.P., Poissant L., Gauchard P.A. & Boutron C.F. 2003. Diurnal cycles of gaseous mercury within the snowpack at Kuujjuarapik/Whapmagoostui, Quebec, Canada. Environmental Science & Technology 37, 3289–3297. [Crossref]

Dommergue A., Sprovieri F., Pirrone N., Ebinghaus R., Brooks S., Courteaud J. & Ferrari C.P. 2010. Overview of mercury measurements in the Antarctic troposphere. Atmospheric Chemistry and Physics 10, 3309–3319. [Crossref]

Donaldson S.G., van Oostdam J., Tikhonov C., Feeley M., Armstrong B., Ayotte P., Boucher O., Bowers W., Chan L., Dallaire F., Dallaire R., Dewailly E., Edwards J., Egeland G.M., Fontaine J., Furgal C., Leech T., Loring E., Muckle G., Nancarrow T., Pereg D., Plusquellec P., Potyrala M., Receveur O. & Shearer R.G. 2010. Environmental contaminants and human health in the Canadian Arctic. Science of the Total Environment 408, 5165–5234. [Crossref]

Douglas T.A., Sturm M., Simpson W.R., Blum J.D., Alvarez-Aviles L., Keeler G.J., Perovich D.K., Biswas A. & Johnson K. 2008. Influence of snow and ice crystal formation and accumulation on mercury deposition to the Arctic. Environmental Science & Technology 42, 1542–1551. [Crossref]

Douglas T.A., Sturm M., Simpson W.R., Brooks S., Lindberg S.E. & Perovich D.K. 2005. Elevated mercury measured in snow and frost flowers near Arctic sea ice leads. Geophysical Research Letters 32, L04502, doi: 10.1029/2004GL022132. [Crossref]

Durnford D., Dastoor A., Figueras-Nieto D. & Ryjkov A. 2010. Long range transport of mercury to the Arctic and across Canada. Atmospheric Chemistry and Physics 10, 6063–6086. [Crossref]

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. Environmental Science & Technology 36, 1238–1244. [Crossref]

Eicken H. 2003. From the microscopic, to the macroscopic, to the regional scale: growth, microstructure and properties of sea ice. In Thomas D.N. & Dickmann G.S. (eds.): Sea ice: an introduction to its physics, chemistry, biology and geology. Pp. 22–81. Oxford: Blackwell Science.

Ekstrom E.B. & Morel F.M. 2008. Cobalt limitation of growth and mercury methylation in sulfate-reducing bacteria. Environmental Science & Technology 42, 93–99. [Crossref]

Ekstrom E.B., Morel F.M. & Benoit J.M. 2003. Mercury methylation independent of the acetyl-coenzyme A pathway in sulfate-reducing bacteria. Applied and Environmental Microbiology 69, 5414–5422. [Crossref]

Finke N., Vandieken V. & Jorgensen B.B. 2007. Acetate, lactate, propionate, and isobutyrate as electron donors for iron and sulfate reduction in Arctic marine sediments, Svalbard. FEMS Microbiology Ecology 59, 10–22. [Crossref]

Fitzgerald W.F., Lamborg C.H. & Hammerschmidt C.R. 2007. Marine biogeochemical cycling of mercury. Chemical Reviews 107, 641–662. [Crossref]

Fitzgerald W.F., Mason R. & Vandal G.M. 1991. Atmospheric cycling and air–water exchange of mercury over mid-continental lacusterine regions. Water, Air, and Soil Pollution 56, 745–767. [Crossref]

Fleming E.J., Mack E.E., Green P.G. & Nelson D.C. 2006. Mercury methylation from unexpected sources: molybdate-inhibited freshwater sediments and an iron-reducing bacterium. Applied and Environmental Microbiology 72, 457–464. [Crossref]

Galand P.E., Casamayor E.O., Kirchman D.L. & Lovejoy C. 2009. Ecology of the rare microbial biosphere of the Arctic Ocean. Proceedings of the National Academy of Sciences of the United States of America 106, 22427–22432. [Crossref]

Galand P.E., Lovejoy C., Pouliot J., Garneau M.-E. & Vincent W.F. 2008. Microbial community diversity and heterotrophic production in a coastal Arctic ecosystem: a stamukhi lake and its source waters. Limnology and Oceanography 53, 813–823. [Crossref]

Garcia E., Amyot M. & Ariya P.A. 2005. Relationship between DOC photochemistry and mercury redox transformations in temperate lakes and wetlands. Geochimica et Cosmochimica Acta 69, 1917–1924. [Crossref]

Garcia E., Poulain A.J., Amyot M. & Ariya P.A. 2005. Diel variations in photoinduced oxidation of Hg(0) in freshwater. Chemosphere 59, 977–981. [Crossref]

Gilmour C.C., Henry E.A. & Mitchell R. 1992. Sulfate stimulation of mercury methylation in freshwater sediments. Environmental Science & Technology 26, 2281–2287. [Crossref]

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. Limonology and Oceanography 47, 967–975. [Crossref]

Gray J.E., Hines M.E., Higueras P.L., Adatto I. & Lasorsa B.K. 2004. Mercury speciation and microbial transformations in mine wastes, stream sediments, and surface waters at the Almaden Mining District, Spain. Environmental Science & Technology 38, 4285–4292. [Crossref]

Hammerschmidt C.R. & Fitzgerald W.F. 2006. Photodecomposition of methylmercury in an Arctic Alaskan lake. Environmental Science & Technology 40, 1212–1216. [Crossref]

Hammerschmidt C.R., Fitzgerald W.F., Lamborg C.H., Balcom P.H. & Tseng C.M. 2006. Biogeochemical cycling of methylmercury in lakes and tundra watersheds of Arctic Alaska. Environmental Science & Technology 40, 1204–1211. [Crossref]

Hayashi K., Kawai S., Ohno T. & Maki Y. 1977. Photomethylation of inorganic mercury by aliphatic α-amino-acids. Journal of the Chemical Society, Chemical Communications 5, 158–159. [Crossref]

Heimann M. & Reichstein M. 2008. Terrestrial ecosystem carbon dynamics and climate feedbacks. Nature 451, 289–292. [Crossref]

Hines M.E., Crill P.M., Varner R.K., Talbot R.W., Shorter J.H., Kolb C.E. & Harriss R.C. 1998. Rapid consumption of low concentrations of methyl bromide by soil bacteria. Applied and Environmental Microbiology 64, 1864–1870.

Hines M.E. & Duddleston K.N. 2001. Carbon flow to acetate and C1 compounds in northern wetlands. Geophysical Research Letters 28, 4251–4254. [Crossref]

Hines M.E., Horvat M., Faganeli J., Bonzongo J.C.J., 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. Environmental Research 83, 129–139. [Crossref]

Jensen S. & Jernelöv A. 1969. Biological methylation of mercury in aquatic organisms. Nature 223, 753–754. [Crossref]

Johansen P., Mulvad G., Pedersen H.S., Hansen J.C. & Riget F. 2007. Human accumulation of mercury in Greenland. Science of the Total Environment 377, 173–178. [Crossref]

Junge K., Eicken H. & Deming J.W. 2004. Bacterial activity at −2 to −20 °C in Arctic wintertime sea ice. Applied and Environmental Microbiology 70, 550–557. [Crossref]

Kawamura K., Yanase A., Eguchi T., Mikami T. & Barrie L.A. 1996. Enhanced atmospheric transport of soil derived organic matter in spring over the High Arctic. Geophysical Research Letters 23, 3735–3738. [Crossref]

Kelly C.A., Rudd J.W.M., Bodaly R.A., Roulet N.P., St. Louis V.L., Heyes A., Moore T.R., Schiff S., Aravena R., Scott K.J., Dyck B., Harris R., Warner B. & Edwards G. 1997. Increases in fluxes of greenhouse gases and methyl mercury following flooding of an experimental reservoir. Enviornmental Science & Technology 31, 1334–1344. [Crossref]

Kelly C.A., Rudd J.W. & Holoka M.H. 2003. Effect of pH on mercury uptake by an aquatic bacterium: implications for Hg cycling. Enviornmental Science & Technology 37, 2941–2946. [Crossref]

Kerin E.J., Gilmour C.C., Roden E., Suzuki M.T., Coates J.D. & Mason R.P. 2006. Mercury methylation by dissimilatory iron-reducing bacteria. Applied and Environmental Microbiology 72, 7919–7921. [Crossref]

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. Applied and Environmental Microbiology 66, 2430–2437. [Crossref]

Kirchman D.L., Elifantz H., Dittel A.I., Malmstrom R.R. & Cottrell M.T. 2007. Standing stocks and activity of Archaea and Bacteria in the western Arctic Ocean. Limnology and Oceanography 52, 495–507. [Crossref]

Kirk J.L., St. Louis V.L., Hintelmann H., Lehnherr I., Else B. & Poissant L. 2008. Methylated mercury species in marine waters of the Canadian High and sub Arctic. Environmental Science & Technology 42, 8367–8373. [Crossref]

Klaminder J., Yoo K., Rydberg J. & Giesler R. 2008. An explorative study of mercury export from a thawing palsa mire. Journal of Geophysical Research—Biogeosciences 113, G04034, doi: 10.1029/2008JG000776. [Crossref]

Knoblauch C., Jorgensen B.B. & Harder J. 1999. Community size and metabolic rates of psychrophilic sulfate-reducing bacteria in Arctic marine sediments. Applied and Environmental Microbiology 65, 4230–4233.

Koh E.Y., Atamna-Ismaeel N., Martin A., Cowie R.O., Beja O., Davy S.K., Maas E.W. & Ryan K.G. 2010. Proteorhodopsin-bearing bacteria in Antarctic sea ice. Applied and Environmental Microbiology 76, 5918–5925. [Crossref]

Krabbenhoft D.P., Hurley J.P., Olson M.L. & Cleckner L.B. 1998. Diel variability of mercury phase and species distributions in the Florida Everglades. Biogeochemistry 40, 311–325. [Crossref]

Krembs C. & Deming J. (2008). The role of exoplymers in microbial adaptation to sea ice. In Margesin R. et al. (eds.): Psychrophiles: from biodiversity to biotechnology. Pp. 247–264. Berlin: Springer.

Kritee K., Barkay T. & Blum J.D. 2009. Mass dependent stable isotope fractionation of mercury during mer mediated microbial degradation of monomethylmercury. Geochimica et Cosmochimica Acta 73, 1285–1296. [Crossref]

Krummel E.M., Gregory-Eaves I., Macdonald R.W., Kimpe L.E., Demers M.J., Smol J.P., Finney B. & Blais J.M. 2005. Concentrations and fluxes of salmon-derived polychlorinated biphenyls (PCBs) in lake sediments. Environmental Science & Technology 39, 7020–7026. [Crossref]

Lalonde J.D., Amyot M., Doyon M.R. & Auclair J.C. 2003. Photo-induced Hg(II) reduction in snow from the remote and temperate Experimental Lakes Area (Ontario, Canada). Journal of Geophysical Research—Atmospheres 108, article no. 4200, doi: 10.1029/200iJD001534. [Crossref]

Lalonde J.D., Amyot M., Kraepiel A.M.L. & Morel F.M.M. 2001. Photooxidation of Hg(0) in artificial and natural waters. Environmental Science & Technology 35, 1367–1372. [Crossref]

Lalonde J.D., Amyot M., Orvoine J., Morel F.M.M., Auclair J.C. & Ariya P.A. 2004. Photoinduced oxidation of Hg0(aq) in the waters from the St. Lawrence estuary. Environmental Science & Technology 38, 508–514. [Crossref]

Lalonde J.D., Poulain A.J. & Amyot M. 2002. The role of mercury redox reactions in snow on snow-to-air mercury transfer. Environmental Science & Technology 36, 174–178. [Crossref]

Larose C., Berger S., Ferrari C., Navarro E., Dommergue A., Schneider D. & Vogel T.M. 2010. Microbial sequences retrieved from environmental samples from seasonal Arctic snow and meltwater from Svalbard, Norway. Extremophiles 14, 205–212. [Crossref]

Larose C., Dommergue A., de Angelis M., Cossa D., Averty B., Marusczak N., Soumis N., Schneider D. & Ferrari C.P. 2010. Springtime changes in snow chemistry lead to new insights into mercury methylation in the Arctic. Geochimica et Cosmochimica Acta 74, 6263–6275. [Crossref]

Lawrence D.M. & Slater A.G. 2005. A projection of severe near-surface permafrost degradation during the 21st century. Geophysical Research Letters 32, L24401, doi: 10.1029/2005GL025080. [Crossref]

Lehnherr I. & St. Louis V.L. 2009. Importance of ultraviolet radiation in the photodemethylation of methylmercury in freshwater ecosystems. Environmental Science & Technology 43, 5692–5698. [Crossref]

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. Environmental Science & Technology 36, 1245–1256. [Crossref]

Loseto L.L., Lean D.R. & Siciliano S.D. 2004. Snowmelt sources of methylmercury to High Arctic ecosystems. Environmental Science & Technology 38, 3004–3010. [Crossref]

Loseto L.L., Siciliano S.D. & Lean D.R. 2004. Methylmercury production in High Arctic wetlands. Environmental Toxicology & Chemistry 23, 17–23. [Crossref]

Loseto L.L., Stern G.A., Deibel D., Connelly T.L., Prokopowicz A., Lean D.R.S., Fortier L. & Ferguson D.R.S. 2008. Linking mercury exposure to habitat and feeding behaviour in Beaufort Sea beluga whales. Journal of Marine Systems 3–4, 1012–1024. [Crossref]

Macdonald R.W., Harner T. & Fyfe J. 2005. Recent climate change in the Arctic and its impact on contaminant pathways and interpretation of temporal trend data. Science of the Total Environment 342, 5–86. [Crossref]

Macdonald R.W. & Loseto L.L. 2010. Are Arctic Ocean ecosystems exceptionally vulnerable to global emissions of mercury? A call for emphasised research on methylation and the consequences of climate change. Environmental Chemistry 7, 133–138. [Crossref]

Malcolm E.G., Schaefer J.K., Ekstrom E.B., Tuit C.B., Jayakumar A., Park H., Ward B.B. & Morel E.M.M. 2010. Mercury methylation in oxygen deficient zones of the oceans: no evidence for the predominance of anaerobes. Marine Chemistry 122, 11–19. [Crossref]

Manganelli M., Malfatti F., Samo T.J., Mitchell B.G., Wang H. & Azam F. 2009. Major role of microbes in carbon fluxes during austral winter in the southern Drake Passage. PLoS One 4, e6941, [Crossref]

Marvin-Dipasquale 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. Environmental Science & Technology 34, 4908–4917. [Crossref]

Marvin-Dipasquale M.C. & Oremland R.S. 1998. Bacterial methylmercury degradation in Florida Everglades peat sediment. Environmental Science & Technology 32, 2556–2563. [Crossref]

Marx J.G., Carpenter S.D. & Deming J.W. 2009. Production of cryoprotectant extracellular polysaccharide substances (EPS) by the marine psychrophilic bacterium Colwellia psychrerythraea strain 34H under extreme conditions. Canadian Journal of Microbiolgoy 55, 63–72. [Crossref]

Mason R.P., Morel F.M.M. & Hemond H.F. 1995. The role of microorganisms in elemental mercury formation in natural-waters. Water, Air, and Soil Pollution 80, 775–787. [Crossref]

Mergler D., Anderson H.A., Chan L.H., Mahaffey K.R., Murray M., Sakamoto M. & Stern A.H. 2007. Methylmercury exposure and health effects in humans: a worldwide concern. Ambio 36, 3–11. [Crossref]

Methe B.A., Nelson K.E., Deming J.W., Momen B., Melamud E., Zhang X.J., Moult J., Madupu R., Nelson W.C., Dodson R.J., Brinkac L.M., Daugherty S.C., Durkin A.S., DeBoy R.T., Kolonay J.F., Sullivan S.A., Zhou L.W., Davidsen T.M., Wu M., Huston A.L., Lewis M., Weaver B., Weidman J.F., Khouri H., Utterback T.R., Feldblyum T.V. & Fraser C.M. 2005. The psychrophilic lifestyle as revealed by the genome sequence of Colwellia psychrerythraea 34H through genomic and proteomic analyses. Proceedings of the National Academy of Sciences of the United States of America 102, 10913–10918. [Crossref]

Miller R.V., Gammon K. & Day M.J. 2009. Antibiotic resistance among bacteria isolated from seawater and penguin fecal samples collected near Palmer Station, Antarctica. Canadian Journal of Microbiology 55, 37–45. [Crossref]

Mindlin S., Minakhin L., Petrova M., Kholodii G., Minakhina S., Gorlenko Z. & Nikiforov V. 2005. Present-day mercury resistance transposons are common in bacteria preserved in permafrost grounds since the Upper Pleistocene. Research in Microbiology 156, 994–1004. [Crossref]

Møller A.K., Barkay T., Abu Al-Soud W., Sørensen S.J., Skov H. & Kroer N. 2011. Diversity and characterization of mercury resistant bacteria in snow, freshwater and sea-ice brine from the High Arctic. FEMS Microbiology Ecology 75, 390–401. [Crossref]

Monperrus M., Tessier E., Amouroux D., Leynaert A., Huonnic P. & Donarda O.F.X. 2007. Mercury methylation, demethylation and reduction rates in coastal and marine surface waters of the Mediterranean Sea. Marine Chemistry 107, 49–63. [Crossref]

Morel F.M.M., Kraepiel A.M.L. & Amyot M. 1998. The chemical cycle and bioaccumulation of mercury. Annual Reviews in Ecology and Systematics 29, 543–566. [Crossref]

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. Applied and Environmental Microbiology 60, 4059–4065.

Niki H., Maker P.S., Savage C.M. & Breitenbach L.P. 1983. A Fourier transform infrared study of the kinetics and mechanism for the reaction Cl+CH3HgCH3. Journal of Physical Chemistry 87, 3722–3724. [Crossref]

O'Driscoll N.J., Lean D.R.S., Loseto L.L., Carignan R. & Siciliano S.D. 2004. Effect of dissolved organic carbon on the photoproduction of dissolved gaseous mercury in lakes: potential impacts of forestry. Environmental Science & Technology 38, 2664–2672. [Crossref]

Oiffer L. & Siciliano S.D. 2009. Methyl mercury production and loss in Arctic soil. Science of the Total Environment 407, 1691–1700. [Crossref]

Osborn A.M., Bruce K.D., Strike P. & Ritchie D.A. 1997. Distribution, diversity and evolution of the bacterial mercury resistance (mer) operon. FEMS Microbiology Reviews 19, 239–262. [Crossref]

Petrova M.A., Mindlin S.Z., Gorlenko Z.M., Kalyaeva E.S., Soina V.S. & Bogdanova E.S. 2002. Mercury-resistant bacteria from permafrost sediments and prospects for their use in comparative studies of mercury resistance determinants. Russian Journal of Genetics 38, 1330–1334. [Crossref]

Pirrone N., Cinnirella S., Feng X., Finkelman R.B., Friedli H.R., Leaner J., Mason R., Mukherjee A.B., Stracher G.B., Streets D.G. & Telmer K. 2010. Global mercury emissions to the atmosphere from anthropogenic and natural sources. Atmospheric Chemistry and Physics 10, 5951–5964. [Crossref]

Poissant L., Zhang H.H., Canario J. & Constant P. 2008. Critical review of mercury fates and contamination in the Arctic tundra ecosystem. Science of the Total Environment 400, 173–211. [Crossref]

Polunin N. & Kelly C.D. 1952. Arctic aerobiology; fungi and bacteria, etc., caught in the air during flights over the geographical North Pole. Nature 170, 314–316. [Crossref]

Pongratz R. & Heumann K.G. 1999. Production of methylated mercury, lead, and cadmium by marine bacteria as a significant natural source for atmospheric heavy metals in polar regions. Chemosphere 39, 89–102. [Crossref]

Poulain A.J., Amyot M., Findlay D., Tel-Or S., Barkay T. & Hintelmann H. 2004. Biological and photchemical production of dissolved gaseous mercury in a boreal lake. Limnology and Oceanography 49, 2265–2275. [Crossref]

Poulain A.J., Garcia E., Amyot M., Campbell P.G.C. & Arlya P.A. 2007. Mercury distribution, partitioning and speciation in coastal vs. inland High Arctic snow. Geochimica et Cosmochimica Acta 71, 3419–3431. [Crossref]

Poulain A.J., Lalonde J.D., Amyot M., Shead J.A., Raofie F. & Ariya P.A. 2004. Redox transformations of mercury in an Arctic snowpack at springtime. Atmospheric Environment 38, 6763–6774. [Crossref]

Poulain A.J., Ni Chadhain S.M., Ariya P.A., Amyot M., Garcia E., Campbell P.G.C., Zylstra G.J. & Barkay T. 2007. Potential for mercury reduction by microbes in the High Arctic. Applied and Environmental Microbiology 73, 2230–2238. [Crossref]

Purdy K.J., Nedwell D.B. & Embley T.M. 2003. Analysis of the sulfate-reducing bacterial and methanogenic archaeal populations in contrasting Antarctic sediments. Applied and Environmental Microbiology 69, 3181–3191. [Crossref]

Ranchou-Peyruse M., Monperrus M., Bridou R., Duran R., Amouroux D., Salvado J.C. & Guyoneaud R. 2009. Overview of mercury methylation capacities among anaerobic bacteria including representatives of the sulphate-reducers: implications for environmental studies. Geomicrobiology Journal 26, 1–8. [Crossref]

Raofie F. & Ariya P.A. 2004. Product study of the gas-phase BrO-initiated oxidation of Hg0: evidence for stable Hg1+ compounds. Environmental Science & Technology 38, 4319–4326. [Crossref]

Ravenschlag K., Sahm K. & Amann R. 2001. Quantitative molecular analysis of the microbial community in marine Arctic sediments (Svalbard). Applied and Environmental Microbiology 67, 387–395. [Crossref]

Rivkina E., Laurinavichius K., McGrath J., Tiedje J., Shcherbakova V. & Gilichinsky D. 2004. Microbial life in permafrost. Advances in Space Research 33, 1215–1221. [Crossref]

Rolfhus K.R. & Fitzgerald W.F. 2004. Mechanisms and temporal variability of dissolved gaseous mercury production in coastal seawater. Marine Chemistry 90, 125–136. [Crossref]

Schaefer J.K., Yagi J., Reinfelder J., Cardona-Marek T., Ellickson K., Tel-Or S. & Barkay T. 2004. The role of the bacterial organomercury lyase (MerB) in controlling methylmercury accumulation in mercury contaminated natural waters. Environmental Science & Technology 34, 4304–4311. [Crossref]

Schroeder W.H., Anlauf K.G., Barrie L.A., Lu J.Y., Steffen A., Schneeberger D.R. & Berg T. 1998. Arctic springtime depletion of mercury. Nature 394, 331–332. [Crossref]

Schroeder W.H. & Munthe J. 1998. Atmospheric mercury—an overview. Atmospheric Environment 32, 809–822. [Crossref]

Scott K.J. 2001. Bioavailable mercury in Arctic snow determined by a light-emitting mer-lux bioreporter. Arctic 54, 92–95.

Segawa T., Miyamoto K., Ushida K., Agata K., Okada N. & Kohshima S. 2005. Seasonal change in bacterial flora and biomass in mountain snow from the Tateyama Mountains, Japan, analyzed by 16S rRNA gene sequencing and real-time PCR. Applied and Environmental Microbiology 71, 123–130. [Crossref]

Selifonova O., Burlage R. & Barkay T. 1993. Bioluminescent sensors for detection of bioavailable Hg(II) in the environment. Applied and Environmental Microbiology 59, 3083–3090.

Selin N.E. 2009. Global biogeochemical cycling of mercury: a review. Annual Review of Envrionment and Resources 34, 43–63. [Crossref]

Sellers P., Kelly C.A., Rudd J.W.M. & MacHutchon A.R. 1996. Photodegradation of methylmercury in lakes. Nature 380, 694–697. [Crossref]

Sherman L.S., Blum J.D., Johnson K.P., Keeler G.J., Barres J.A. & Douglas T.A. 2010. Mass-independent fractionation of mercury isotopes in Arctic snow driven by sunlight. Nature Geoscience 3, 173–177. [Crossref]

Sheu G.R. & Mason R.P. 2004. An examination of the oxidation of elemental mercury in the presence of halide surfaces. Journal of Atmospheric Chemistry 48, 107–130. [Crossref]

Siciliano S.D., O'Driscoll N.J. & Lean D.R. 2002. Microbial reduction and oxidation of mercury in freshwater lakes. Environmental Science & Technology 36, 3064–3068. [Crossref]

Siciliano S.D., O'Driscoll N.J., Tordon R., Hill J., Beauchamp S. & Lean D.R. 2005. Abiotic production of methylmercury by solar radiation. Environmental Science & Technology 39, 1071–1077. [Crossref]

Simbahan J., Kurth E., Schelert J., Dillman A., Moriyama E., Jovanovich S. & Blum P. 2005. Community analysis of a mercury hot spring supports occurrence of domain-specific forms of mercuric reductase. Applied and Environmental Microbiology 71, 8836–8845. [Crossref]

Skov H., Brooks S., Goodsite M.E., Lindberg S.E., Meyers T.P., Landis M., Larsen M.R.B., Jensen B., McConville G., Chung K.H. & Christensen J. 2006. The fluxes of reactive gaseous mercury measured with a newly developed method using relaxed eddy accumulation. Atmospheric Environment 40, 5452–5463. [Crossref]

Slater F.R., Bruce K.D., Ellis R.J., Lilley A.K. & Turner S.L. 2008. Heterogeneous selection in a spatially structured environment affects fitness tradeoffs of plasmid carriage in pseudomonads. Applied and Environmental Microbiology 74, 3189–3197. [Crossref]

Slater F.R., Bruce K.D., Ellis R.J., Lilley A.K. & Turner S.L. 2010. Determining the effects of a spatially heterogeneous selection pressure on bacterial population structure at the sub-millimetre scale. Microbial Ecology 60, 873–884. [Crossref]

Smith T., Pitts K., McGarvey J.A. & Summers A.O. 1998. Bacterial oxidation of mercury metal vapor, Hg(0). Applied and Environmental Microbiology 64, 1328–1332.

St. Louis V.L., Hintelmann H., Graydon J.A., Kirk J.L., Barker J., Dimock B., Sharp M.J. & Lehnherr I. 2007. Methylated mercury species in Canadian High Arctic marine surface waters and snowpacks. Environmental Science & Technology 41, 6433–6441. [Crossref]

St. Louis V.L., Rudd J.W., Kelly C.A., Bodaly R.A., Paterson M.J., Beaty K.G., Hesslein R.H., Heyes A. & Majewski A.R. 2004. The rise and fall of mercury methylation in an experimental reservoir. Environmental Science & Technology 38, 1348–1358. [Crossref]

St. Louis V.L., Sharp M.J., Steffen A., May A., Barker J., Kirk J.L., Kelly D.J., Arnott S.E., Keatley B. & Smol J.P. 2005. Some sources and sinks of monomethyl and inorganic mercury on Ellesmere Island in the Canadian High Arctic. Environmental Science & Technology 39, 2686–2701. [Crossref]

Steffen A., Douglas T., Amyot M., Ariya P., Aspmo K., Berg T., Bottenheim J., Brooks S., Cobbett F., Dastoor A., Dommergue A., Ebinghaus R., Ferrari C., Gardfeldt K., Goodsite M.E., Lean D., Poulain A., Scherz C., Skov H., Sommar J. & Temme T. 2007. A synthesis of atmospheric mercury depletion event chemistry linking atmosphere, snow and water. Atmospheric Chemistry and Physics Discussions 7, 10837–10931. [Crossref]

Steffen A., Schroeder W., Bottenheim J., Narayan J. & Fuentes J.D. 2002. Atmospheric mercury concentrations: measurements and profiles near snow and ice surfaces in the Canadian Arctic during Alert 2000. Atmospheric Environment 36, 2653–2661. [Crossref]

Streets D.G., Zhang Q. & Wu Y. 2009. Projections of global mercury emissions in 2050. Environmental Science & Technology 43, 2983–2988. [Crossref]

Summers A.O. 1992. Untwist and shout: a heavy metal-responsive transcriptional regulator. Journal of Bacteriology 174, 3097–3101.

Sunderland E.M., Krabbenhoft D.P., Moreau J.W., Strode S.A. & Landing W.M. 2009. Mercury sources, distribution, and bioavailability in the North Pacific Ocean: insights from data and models. Global Biogeochemical Cycles 23, GB2010, doi: 10.1029/2008GB003425. [Crossref]

Teitzel G.M. & Parsek M.R. 2003. Heavy metal resistance of biofilm and planktonic Pseudomonas aeruginosa. Applied and Environmental Microbiology 69, 2313–2320. [Crossref]

Van Oostdam J., Donaldson S.G., Feeley M., Arnold D., Ayotte P., Bondy G., Chan L., Dewaily E., Furgal C.M., Kuhnlein H., Loring E., Muckle G., Myles E., Receveur O., Tracy B., Gill U. & Kalhok S. 2005. Human health implications of environmental contaminants in Arctic Canada: a review. Science of the Total Environment, 351/352, 165–246. [Crossref]

Vlassov A.V., Kazakov S.A., Johnston B.H. & Landweber L.F. 2005. The RNA world on ice: a new scenario for the emergence of RNA information. Journal of Molecular Evolution 61, 264–273. [Crossref]

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. [Crossref]

Weber J.H. 1993. Review of possible paths for abiotic methylation of mercury(II) in the aquatic environment. Chemosphere 26, 2063–2077. [Crossref]

Westöö G. 1966. Determination of methylmercury compounds in foodstuffs. I. Methylmercury compounds in fish, identification and determination. Acta Chemica Scandinavia 20, 2131–2137. [Crossref]

Whalin L.M. & Mason R.P. 2006. A new method for the investigation of mercury redox chemistry in natural waters utilizing deflatable Teflon (R) bags and additions of isotopically labeled mercury. Analytica Chimica Acta 558, 211–221. [Crossref]

Wiatrowski H.A., Das S., Kukkadapu R., Ilton E.S., Barkay T. & Yee N. 2009. Reduction of Hg(II) to Hg(O) by magnetite. Environmental Science & Technology 43, 5307–5313. [Crossref]

Wiatrowski H.A., Ward P.M. & Barkay T. 2006. Novel reduction of mercury(II) by mercury-sensitive dissimilatory metal reducing bacteria. Environmental Science & Technology 40, 6690–6696. [Crossref]

Wren C.D. 1986. A review of metal accumulation and toxicity in wild mammals.1. Mercury. Environmental Research 40, 210–244. [Crossref]

Yergeau E., Arbour M., Brousseau R., Juck D., Lawrence J.R., Masson L., Whyte L.G. & Greer C.W. 2009. Microarray and real-time PCR analyses of the responses of High-Arctic soil bacteria to hydrocarbon pollution and bioremediation treatments. Applied and Environmental Microbiology 75, 6258–6267. [Crossref]

Yu R.-Q., Adatto I., Montesdeoca M.R., Driscoll C.T., Hines M.E. & Barkay T. 2010. Mercury methylation in sphagnum moss mats and its association with sulfate-reducing bacteria in an acidic Adirondack forest lake wetland. FEMS Microbiology Ecology 74, 655–668. [Crossref]

Zhang H., Dill C., Kuiken T., Ensor M. & Crocker W.C. 2006. Change of dissolved gaseous mercury concentrations in a southern reservoir lake (Tennessee) following seasonal variation of solar radiation. Environmental Science & Technology 40, 2114–2119. [Crossref]

Zhang T. & Hsu-Kim H. 2010. Photolytic degradation of methylmercury enhanced by binding to natural organic ligands. Nature Geoscience 3, 473–476. [Crossref]