Tải bản đầy đủ (.pdf) (14 trang)

Mercury Hazards to Living Organisms - Chapter 3 pptx

Bạn đang xem bản rút gọn của tài liệu. Xem và tải ngay bản đầy đủ của tài liệu tại đây (194.79 KB, 14 trang )

23
C
HAPTER
3
Properties
Mercury, a silver-white metal that is liquid at room temperature and highly volatile, can exist in
three oxidation states: elemental mercury (Hg
o
), mercurous ion (Hg
2
2+
), and mercuric ion (Hg
2+
).
It can be part of both inorganic and organic compounds (USEPA, 1980; Clarkson et al., 1984). All
mercury compounds interfere with thiol metabolism, causing inhibition or inactivation of proteins
containing thiol ligands and ultimately to mitotic disturbances (Das et al., 1982; Elhassani, 1983;
Hook and Hewitt, 1986). The mercuric species is the most toxic inorganic chemical form, but all
three forms of inorganic mercury may have a common molecular mechanism of damage in which
Hg
2+
is the toxic species (Clarkson and Marsh, 1982). Mercury poisoning and treatment is discussed
in more detail in Chapter 4.
3.1 PHYSICAL
Pure mercury is a coherent, silvery-white mobile liquid with a metallic luster (Anon., 1948). In
thin layers it transmits a bluish-violet light. It freezes at about minus 39
°C with contraction, forming
a white, ductile, malleable mass easily cut with a knife, and with cubic crystals. When heated, the
metal expands uniformly, boiling at 357.01°C at 760 mm, and vaporizing at about 360.0
°
C. The


vapor is colorless. Mercury forms two well-defined series of salts: the mercurous salts derived from
the oxide Hg
2
O, and the mercuric salts from the oxide HgO. Mercuric oxide occurs in two forms:
a bright red crystalline powder and as an orange-yellow powder. The yellow form is the most
reactive and is transformed into the red when heated at 400.0
°
C. Heating the red form results in a
black compound, which regains its color on cooling; on further heating to 630.0
°
C, it decomposes
to mercury and oxygen (Anon., 1948).
Mercurous and mercuric chloride, known respectively as calomel and corrosive sublimate, are
two of the most important salts of mercury (Table 3.1). Other halogenated mercury salts include
mercurous bromide, Hg
2
Br
2
, a yellowish-white powder insoluble in water; mercuric bromide,
HgBr
2
, comprised of white crystals sparingly soluble in cold water but readily soluble in hot water;
mercurous iodide, Hg
2
I
2
, a yellowish-green powder; and mercuric iodide, which exists in two
crystalline forms. Other mercuric and mercurous compounds include nitrates, nitrites, sulfides,
sulfates, phosphides, phosphates, and ammonium salts (Anon., 1948).
3.2 CHEMICAL

Elemental mercury is relatively inert in dry air, oxygen, nitrous oxide, carbon dioxide, ammonia,
and some other gases at room temperatures (Anon., 1948). In damp air, it slowly becomes coated
© 2006 by Taylor & Francis Group, LLC
24 MERCURY HAZARDS TO LIVING ORGANISMS
with a film of mercurous oxide. When heated in air or oxygen, it is transformed into the red mercuric
oxide, which decomposes into mercury and oxygen on continued heating at higher temperatures.
Mercury dissolves many metals to form compounds called amalgams (Anon., 1948).
Chemical speciation is probably the most important variable influencing mercury toxicity, but
mercury speciation is difficult to quantify, especially in natural environments (Boudou and Ribeyre,
1983). Mercury compounds in an aqueous solution are chemically complex. Depending on pH,
alkalinity, redox, and other variables, a wide variety of chemical species are liable to be formed,
having different electrical charges and solubilities. For example, HgCl
2
in solution can speciate
into Hg(OH)
2
, Hg
2+
, HgCl
+
, Hg(OH)
-
, HgCl
3
-
, and HgCl
4
2-
; anionic forms predominate in saline
environments (Boudou and Ribeyre, 1983). In the aquatic environment, under naturally occurring

conditions of pH and temperature, mercury may also become methylated by biological or chemical
processes, or both (Beijer and Jernelov, 1979; USEPA, 1980; Ramamoorthy and Blumhagen, 1984;
Zillioux et al., 1993; Figure 3.1) — although biological methylation is limited (Callister and
Winfrey, 1986). Methylmercury is the most hazardous mercury species due to its high stability, its
lipid solubility, and its possession of ionic properties that lead to a high ability to penetrate
membranes in living organisms (Beijer and Jernelov, 1979; Hamasaki et al., 1995). In general,
essentially all mercury in freshwater fish tissues is in the form of methylmercury; however, meth-
ylmercury accounts for less than 1.0% of the total mercury pool in a lake (Regnell, 1990).
All mercury discharged into rivers, bays, or estuaries as elemental (metallic) mercury, inorganic
divalent mercury, or phenylmercury or alkoxyalkyl mercury can be converted into methylmercury
compounds by natural processes (Jernelov, 1969; Figure 3.1). Mercury methylation in ecosystems
depends on mercury loadings, microbial activity, nutrient content, pH and redox conditions, sus-
pended sediment load, sedimentation rates, and other variables (USNAS, 1978; Compeau and
Bartha, 1984; Berman and Bartha, 1986; Callister and Winfrey, 1986; Jackson, 1986). Net meth-
ylmercury production was about 10 times higher in reduced sediments than in oxidized sediments
(Regnell, 1990). The finding that certain microorganisms are able to convert inorganic and organic
forms of mercury into the highly toxic methylmercury or dimethylmercury has made it clear that
any form of mercury is highly hazardous to the environment (USEPA, 1980, 1985). The synthesis
of methylmercury by bacteria from inorganic mercury compounds present in the water or in the
sediments is the major source of this molecule in aquatic environments (Boudou and Ribeyre, 1983;
Nakamura, 1994). This process can occur under both aerobic and anaerobic conditions (Beijer and
Jernelov, 1979; Clarkson et al., 1984), but seems to favor anaerobic conditions (Olson and Cooper,
1976; Callister and Winfrey, 1986). Transformation of inorganic mercury to an organic form by
© 2006 by Taylor & Francis Group, LLC
Table 3.1 Some Properties of Mercury and Its Compounds
a
Property
Elemental
Mercury
Mercurous

Chloride
Mercuric
Chloride
Methylmercury
Chloride
Empirical formula Hg Hg
2
Cl
2
HgCl
2
CH
3
HgCl
Molecular weight 200.59 472.09 271.52 251.09
c
Chlorine, % 0 15.02 26.12 14.12
c
Mercury, % 100 84.98 73.88 79.89
c
Melting point,
°
C
−38.87 Sublimes
at 400–500
277 170
c
Density 13.534 7.15 5.4 4.063
c
Solubility, mg/L (ppm)

in water 0.056 2.0 74,070 1,016
d
in benzene 2.387
b
Insoluble 5,000 6,535
e
a
All data from Merck Index (1976), except where indicated.
b
Spencer and Voigt (1961).
c
Weast and Astle (1982).
d
Eisler (unpublished), 72-h equilibrium value.
e
Eisler (unpublished), 24-h equilibrium value.
PROPERTIES 25
bacteria alters its biochemical reactivity and hence its fate (Windom and Kendall, 1979; Figure 3.1).
Methylmercury is decomposed by bacteria in two phases. First, hydrolytic enzymes cleave the
C–Hg bond, releasing the methyl group. Second, a reductase enzyme converts the ionic mercury
to the elemental form, which is then free to diffuse from the aquatic environment into the vapor
phase. These demethylating microbes appear to be widespread in the environment; they have been
isolated from water, sediments, and soils and from the gastrointestinal tract of mammals, including
humans (Clarkson et al., 1984). Some strains of microorganisms contain mercuric reductase, which
transforms inorganic mercury to elemental mercury, and organomercurial lyase, which degrades
organomercurials to elemental mercury (Baldi et al., 1991).
Humic substances can reduce inorganic divalent mercury (Hg
2+
) to elemental mercury (Hg
o

).
In aquatic environments, Hg
o
was highest under anoxic conditions, in the absence of chloride, and
at pH 4.5 (Allard and Arsenie, 1991). Under these conditions, about 25.0% of 400.0
µg Hg
2+
/L
was reduced to Hg
o
in 50 h. Production of Hg
o
was reduced in the presence of europium ions and
WATER AND BENTHIC REGION
AIR
ELEMENTAL
SHELLFISH
MERCURIC
MERCUROUS
METHYL
DIMETHYL
inorganic and organic complexes
FISH
CH
4
Hg
2+
Hg
0
Hg

0
Hg
2+
Hg
2+
2
HgS
(CH
3
)
2
Hg
(CH
3
)
2
Hg
CH
3
SHgCH
3
CH
3
Hg
+
C
2
H
6
CH

3
Hg
+
© 2006 by Taylor & Francis Group, LLC
Figure 3.1 Major transformations of mercury in the environment. (Modified from Beijer and Jernelov, 1979;
Nakamura, 1994; and Eisler, 2000.)
26 MERCURY HAZARDS TO LIVING ORGANISMS
by methylated carboxyl groups in the humic substances (Allard and Arsenie, 1991). Mercury is
efficiently transferred through wetlands and forests in a more reactive form relative to other land
use patterns, resulting in an increased uptake by organisms inhabiting these rivers or downstream
impoundments and drainage lakes (Hurley et al., 1995). The behavior and accumulation of mercury
in forest soils of Guyana, South America, is related to the penetration of humic substances and the
progressive adsorption onto iron oxy-hydroxides in the mineral horizons; flooding of these soils
may lead to a release of 20.0%
of the mercury initially present (Roulet and Lucotte, 1995).
Methylmercury is produced by methylation of inorganic mercury present in both freshwater
and saltwater sediments, and accumulates in aquatic food chains in which the top-level predators
usually contain the highest concentrations (Clarkson and Marsh, 1982). The percent of total mercury
accounted for by methylmercury generally increases with higher trophic levels, confirming that
methylmercury is more efficiently transferred to higher trophic levels than inorganic mercury
compounds (Becker and Bigham, 1995). Organomercury compounds other than methylmercury
decompose rapidly in the environment, and behave much like inorganic mercury compounds (Beijer
and Jernelov, 1979). In organisms near the top of the food chain, such as carnivorous fishes, almost
all mercury accumulated is in the methylated form, primarily as a result of the consumption of
prey containing methylmercury; methylation also occurs at the organism level by way of mucus,
intestinal bacteria, and enzymatic processes, but these pathways are not as important as diet
(Huckabee et al., 1979; Boudou and Ribeyre, 1983). In tissues of marine flounders, inorganic
mercury compounds are strongly bound to metallothioneins and high-molecular-weight ligands;
however, methylmercury has a low affinity for metallothioneins and is strongly lipophilic (Barghi-
giani et al., 1989).

3.3 BIOLOGICAL
The biological cycle of mercury is delicately balanced, and small changes in input rates, geophysical
conditions, and the chemical form of mercury can result in increased methylation rates in sensitive
systems (USNAS, 1978). For example, the acidification of natural bodies of freshwater is statisti-
cally associated with elevated concentrations of methylmercury in the edible tissues of predatory
fishes (Clarkson et al., 1984). Acidification has a stronger effect on the supply of methylmercury
to the ecosystem than on specific rates of uptake by the biota (Bloom et al., 1991). In chemically
sensitive waterways, such as poorly buffered lakes, the combined effects of acid precipitation and
increased emissions of mercury to the atmosphere (with subsequent deposition) pose a serious
threat to the biota if optimal biomethylation conditions are met (USNAS, 1978). In remote lakes
of the Adirondack mountain region in upstate New York, fish contain elevated mercury concentra-
tions in muscle; mercury loadings in fish were directly associated with decreasing water column
pH and increasing concentrations of dissolved organic carbon (DOC), although high DOC concen-
trations may complex methylmercury, thus diminishing its bioavailability (Driscoll et al., 1995).
At high concentrations of monomeric aluminum, the complexation of methylmercury with DOC
decreases, enhancing the bioavailability of methylmercury (Driscoll et al., 1995).
Mercury excretion in mammals is through the urine and feces, and depends on the form of
mercury, dose, and time postexposure (Goyer, 1986). With mercury vapor, there is minor initial
loss from exhalation and major loss via fecal excretion. With inorganic mercury, fecal loss is
predominant after initial exposure, and renal excretion increases with time. With methylmercury,
about 90.0%
is excreted in feces after acute or chronic exposure and does not change over time
(Goyer, 1986).
Based on animal studies, all forms of mercury can cross the placenta to the fetus (Goyer, 1986).
Fetal uptake of elemental mercury in rats — possibly due to its high solubility in lipids — is 10
to 40 times higher than uptake after exposure to inorganic mercury salts. Concentrations of mercury
in the fetus after exposure to alkylmercuric compounds, when compared to elemental mercury, are
© 2006 by Taylor & Francis Group, LLC
PROPERTIES 27
twice those found in maternal tissues, and methylmercury levels in fetal red blood cells are 30.0%

higher than in maternal cells. The positive fetal maternal gradient and increased mercury concen-
tration in fetal erythrocytes enhance fetal toxicity to mercury, especially after exposure to alkyl-
mercury. Maternal milk contains about 5.0%
of the mercury concentration of maternal blood;
however, neonatal exposure to mercury may be greatly augmented by nursing (Goyer, 1986).
Elemental or metallic mercury is oxidized, probably via catalases, to divalent mercury after
absorption into body tissues (Goyer, 1986). Most inhaled mercury vapor absorbed into erythrocytes
is transformed into divalent mercury, but some is also transported as metallic mercury to the brain,
where biotransformation may occur. Some metallic mercury may cross the placenta into the fetus.
Oxidized metallic mercury is then accumulated by brain and fetus. Organomercurials also undergo
biotransformation to divalent mercury compounds in tissues by cleavage of the carbon–mercury
bond, with no evidence of organomercury formation by mammalian tissues. Phenylmercurials are
converted to inorganic mercury more rapidly than the shorter-chain methylmercurials. Phenyl and
methoxyethyl mercurials are excreted at about the same rate as inorganic mercury, whereas methyl-
mercury excretion is slower (Goyer, 1986). The half-time persistence of methylmercurials in
mammalian tissues is about 70 days and seems to follow a linear pattern. For inorganic mercury
salts, the biological half-time is about 40 days. And for elemental mercury or mercury vapor, the
half-time persistence in tissues ranges between 35 and 90 days and also appears to be linear (Goyer,
1986).
3.4 BIOCHEMICAL
Mercury binds strongly with sulfhydryl groups, and has many potential target sites during embryo-
genesis; phenylmercury and methylmercury compounds are among the strongest known inhibitors
of cell division (Birge et al., 1979). In mammalian hepatocytes, the L-alanine carrier contains a
sulfhydryl group that is essential for its activity and is inhibited by mercurials (Sellinger et al.,
1991). In the little skate (Raja erinacea), HgCl
2
inhibits Na
+
-dependent alanine uptake and Na
+

/K
+
-
ATPase activity, and increases K
+
permeability. Inhibition of Na
+
-dependent alanine in skate hepa-
tocytes by HgCl
2
is attributed to three different concentration-dependent mechanisms: (1) direct
interaction with the transporters; (2) dissipation of the Na
+
gradient; and (3) loss of membrane
integrity (Sellinger et al., 1991). Organomercury compounds, especially methylmercury, cross pla-
cental barriers and can enter mammals by way of the respiratory tract, gastrointestinal (GI) tract,
skin, or mucus membranes (Elhassani, 1983). When compared with inorganic mercury compounds,
organomercurials are more completely absorbed, are more soluble in organic solvents and lipids,
pass more readily through biological membranes, and are slower to be excreted (Clarkson and
Marsh, 1982; Elhassani, 1983; Greener and Kochen, 1983). Biological membranes, including those
at the blood–brain interface and placenta, tend to discriminate against ionic and inorganic mercury,
but allow relatively easy passage of methylmercury and dissolved mercury vapor (Greener and
Kochen, 1983). As judged by membrane model studies, it appears that electrically neutral mercurials
are responsible for most of the diffusion transport of mercury, although this movement is modified
significantly by pH and mercury speciation. It appears, however, that the liposolubility of meth-
ylmercury is not the entire reason for its toxicity and does not play a major role in its transport.
This hypothesis should be examined further in studies with living membranes (Boudou et al., 1983).
In liver cells, methylmercury forms soluble complexes with cysteine and glutathione, which
are secreted in bile and reabsorbed from the GI tract. In general, however, organomercurials undergo
cleavage of the carbon–mercury bond, releasing ionic inorganic mercury (Goyer, 1986). Mercuric

mercury induces synthesis of metallothioneins, mainly in kidney cells. Mercury within renal cells
becomes localized in lysosomes (Goyer, 1986).
The finding of naturally elevated mercury concentrations in marine products of commerce is
worrisome to regulatory agencies charged with protection of human health. This topic is discussed
© 2006 by Taylor & Francis Group, LLC
28 MERCURY HAZARDS TO LIVING ORGANISMS
later in Chapter 6 and Chapter 12. At this time, however, many authorities argue that mercury–sele-
nium interactions are the key to methylmercury bioavailability in human diets. In the case of marine
mammals and seabirds, for example, mercury is accumulated from the diet mainly as methylmercury
and then transformed into the less toxic inorganic mercury. Accordingly, most of the tissue mercury
found in high concentrations of these two marine groups is inorganic mercury (Itano et al., 1984;
Thompson, 1990; Law, 1996; Yang et al., 2002). Many authorities aver that selenium detoxifies
inorganic mercury by forming complexes in a 1:1 molar ratio (Koeman et al., 1973; Ping et al.,
1986; Palmisano et al., 1995). Field studies on marine mammal livers corroborate the equimolar
ratio of mercury and selenium (Koeman et al., 1973; Pelletier, 1985; Cuvin-Aralar and Furness,
1991; Kim et al., 1996); moreover, mercuric selenide (HgSe) — an inert end product of mercury
detoxification in marine mammals — was found in the livers of marine mammals and birds (Martoja
and Berry, 1980; Rawson et al., 1995; Nigro and Leonzio, 1996). A proposed mercury detoxification
model in higher-trophic marine animals involves transformation of dietary methylmercury to inor-
ganic mercury by reactive oxygen species (Suda et al., 1992; Hirayama and Yasutake, 2001;
Yasutake and Hirayama, 2001); gut microflora (Rowland et al., 1984; Rowland, 1988); and selenium
(Iwata et al., 1982). Inorganic mercury binds to metallothioneins or forms an equimolar Hg–Se
complex, and subsequently combines with high molecular weight compounds in liver (Ikemoto
et al., 2004). Glutathione molecules solubilize HgSe; the complex then binds to a specific protein
(Gailer et al., 2000). The protein-bound Hg–Se complex is thought to be the precursor of mercuric
selenide, which occurs in macrophages of marine mammals (Nigro, 1994; Rawson et al., 1995;
Nigro and Leonzio, 1996).
3.5 MERCURY TRANSPORT AND SPECIATION
The global mercury cycle involves mercury release from geological and industrial processes into
water and the atmosphere, followed by sedimentation via rainfall and by microbial metabolism that

releases mercury from soil and sediments and transforms mercury from one chemical form to another
(WHO, 1976, 1989, 1990, 1991; Fitzgerald, 1986; Fitzgerald and Clarkson, 1991; Barkay, 1992).
Atmospheric-borne mercury, including anthropogenic mercury, is deposited everywhere, includ-
ing remote areas of the globe, hundreds of kilometers from the nearest mercury source, as evidenced
by its presence in ancient lake sediments (Fitzgerald et al., 1998) and glacial ice (Schuster et al.,
2002). In Amituk Lake in the Canadian Arctic, recent annual deposition of mercury was estimated
at 15.1 kg, about 56.0%
from snowpack and the rest from precipitation (Semkin et al., 2005). This
represents a dramatic increase from historic annual burdens of 6.0 kg mercury annually in this
remote area (Semkin et al., 2005); the effects of this increase on Arctic watersheds is unknown.
Many important sources affecting global mercury cycles emit elemental metallic mercury (Hg
o
)
in gaseous form, and to a lesser extent gaseous and particulate species of Hg
2+
(USEPA, 1997;
Pacyna et al., 2001). Gaseous and particulate Hg
2+
are removed from the atmosphere through rainfall
and dry deposition, thus limiting long-range transport (Lindberg and Stratton, 1998; Ebinghaus
et al., 2002). Inorganic Hg
2+
can be readily reduced to Hg
o
by natural processes in terrestrial and
aquatic ecosystems (Nakamura, 1994; Zhang and Lindberg, 2001). Elemental mercury can be
oxidized in the atmosphere to Hg
2+
, which is removed through wet and dry deposition (Lindberg
et al., 2002). About 67.0%

of the mercury in global fluxes is a result of human activities and the
rest from natural emissions (Mason et al., 1994). Soils and sediments are the primary sinks for
atmospherically derived mercury; however, these enriched pools can be remobilized through vol-
atilization, leaching, and erosion (Wiener et al., 2003).
Mercury speciation varies in atmospheric, aquatic, and terrestrial environments. In the atmo-
sphere, mercury is in the form of gaseous elemental Hg
o
(95.0%), Hg
2+
(called reactive gaseous
© 2006 by Taylor & Francis Group, LLC
PROPERTIES 29
mercury), and trace amounts of methylmercury (Lindberg and Stratton, 1998; Schroeder and
Munthe, 1998). Particulate and reactive mercury in the atmosphere travels short distances, usually
less than 50 km, and has a residence time of about 1 year (Mason et al., 1994). Reactive gaseous
mercury is assumed to be HgCl
2
, with some Hg(NO
3
)
2
•H
2
O in the gas phase (Stratton et al., 2001).
Reactive gaseous mercury may comprise a majority of atmospheric gaseous mercury at some
locations (e.g., springtime in Alaska) and this component is rapidly removed from the atmosphere
by wet and dry deposition (Lindberg et al., 2002;
Munthe et al., 2001) and available for methylation
once deposited (Lindberg et al., 2002). Mercury point sources and rates of particle scavenging are
key factors in atmospheric transport rates to sites of methylation and subsequent entry into the

marine food chain (Rolfhus and Fitzgerald, 1995). Airborne soot particles transport mercury into
the marine environment either as nuclei for raindrop formation or by direct deposition on water
(Rawson et al., 1995). In early 1990, both dimethylmercury and monomethylmercury species were
found in the subthermocline waters of the equatorial Pacific Ocean; the formation of these alkyl-
mercury species in the low oxygen zone suggests that Hg
2+
is the most likely substrate (Mason and
Fitzgerald, 1991).
In aquatic environments, Hg
o
and methylmercury species are the most common, with concen-
trations low, usually in the picogram-per-liter to microgram-per-liter range, except in the vicinity
of anthropogenic or natural mercury sources (Wiener et al., 2003). The speciation of mercury in
water is influenced by redox, pH, and ligands (Gill and Bruland, 1990). In most aerated surface
waters near pH 7.0, ion-pair formation for CH
3
+
and Hg
2+
is dominated by dissolved organic matter
and chloride (Ravichandran et al., 1999). Under anoxic conditions, Hg
2+
and CH
3
+
are present
mainly as sulfide and sulfhydryl ion pairs (Benoit et al., 1999a; Gill et al., 1999). Complexed Hg
2+
sulfides are less available for methylation (Benoit et al., 1999b). Concentrations of total mercury
in uncontaminated, unfiltered freshwater samples range from 0.3 to 8.0 ng/L, but range from 10.0

to 40.0 ng/L near mercury sources (Wiener et al., 2003), and up to 1000.0 ng/L in waters contam-
inated by mercury tailings from gold mines (Eisler, 2004).
Sulfate-reducing bacteria are the most important mercury-methylating agents in aquatic envi-
ronments (Gilmour et al., 1992, 1998), with the most important site of methylation at the
oxic–anoxic interface in sediments (Pak and Bartha, 1998); a similar pattern is documented for
wetlands (Krabbenhoft et al., 1995; Branfireum et al., 1996; Gilmour et al., 1998). In sediments,
microbial methylation of mercury is fastest in the upper profiles where the rate of sulfate reduction
is greatest (Hines et al., 2000; King et al., 2001). Methylcobalamin, produced by bacteria, is the
active methyl donor to the Hg
2+
ion; methylcobalamin reacts with Hg
2+
to form CH
3
Hg
+
(Choi
et al., 1994). Methylation also occurs, but to a lesser degree, in aerobic freshwater and seawater,
in aquatic plants, and in mucosal slime and intestines of fish (Wiener et al., 2003). Abiotic formation
of CH
3
Hg
+
compounds in sediments is documented; however, amounts formed are small when
compared to biotic processes (Nagase et al., 1988; Falter and Wilken, 1998). Demethylation occurs
via abiotic and biotic processes in the near-surface sediments and in the water column (Winfrey
and Rudd, 1990; Oremland et al., 1991; Sellers et al., 1996). An oxidative demethylation pathway
similar to that of the degradation of methanol or monomethylamine by methanogens has been
proposed for methylmercury degradation (Marvin-DiPasquale et al., 1998, 2000). Photodegradation
of methylmercury in surface waters of freshwater lakes is documented at rates up to 18.0% daily

and is quantitatively important in mercury budgets of that ecosystem (Sellers et al., 1996, 2001;
Branfireun et al., 1998; Krabbenhoft et al., 2002); the end products of mercury photodemethylation
are not known with certainty (Krabbenhoft et al., 1998). Frequently, however, methylmercury
concentrations in aerobic lakewater surfaces increase during sunlight hours or remain unchanged
(Sicialiano et al., 2005). This phenomenon is linked to dissolved organic matter (DOM) and solar
radiation — specifically to certain fractions of DOM that generates methylmercury when exposed
to sunlight, especially in water from lakes with logged watersheds. The mechanism to account for
© 2006 by Taylor & Francis Group, LLC
30 MERCURY HAZARDS TO LIVING ORGANISMS
methylmercury production is not clear at present; however, it corrects the conventional wisdom
that methylmercury is rapidly photodegraded (Sicialiano et al., 2005).
Terrestrial soils are a significant contributor of mercury to surface waters (Mason et al., 1994;
Cooper and Gillespie, 2001; Grigal, 2002; Gabriel and Williamson, 2004). Moreover, up to 60.0%
of the atmospherically deposited mercury that reaches lakes originates from the associated terrestrial
watershed (Krabbenhoft and Babiarz, 1992; Lorey and Driscoll, 1999; Grigal, 2002). The main
biogeochemical reactions affecting the transport and speciation of mercury in the terrestrial water-
shed include formation of mercury ligands, mercury adsorption and desorption, and elemental
mercury reduction and volatilization (Gabriel and Williamson, 2004). In terrestrial environments,
OH

, Cl

, and S

ions have the greatest impact on inorganic mercury–ligand formation. Under
oxidized surface soil conditions, Hg(OH)
2
, HgCl
2
, HgOH

+
, HgS (cinnabar), and Hg
0
are the
dominant inorganic mercury forms and usually bind to organic and mineral ions. Under reducing
conditions, common mercury forms are HgSH
+
, HgOHSH, and HgClSH, and many are further
bound to both inorganic and organic ligands. The following organomercurials predominated in
terrestrial soils: CH
3
HgCl > CH
3
HgOH > free CH
3
Hg
+
(Gabriel and Williamson, 2004).
In upland soils, mercury is mostly in the form of Hg
2+
species sorbed to organic matter in the
humus layer and a lesser extent to soil minerals (Kim et al., 1997). The overall adsorption of
mercury to mineral and organic particles is positively correlated, in order of importance, with
surface area, organic content, cation exchange capacity, and grain size (Gabriel and Williamson,
2004). Mercury adsorption and desorption to mineral and organic surfaces is strongly influenced
by pH and dissolved ions; for example, increased Cl
−−
−−
and decreased pH — alone or together —
can decrease mercury adsorption, and clays and organic soils have the highest capability of

adsorbing mercury (Gabriel and Williamson, 2004). Common forms of methylated mercury in soils
depend on pH (Reimers and Krenkel, 1974; Jackson, 1998). At pH 2 to about 4.7, the most common
forms were CH
3
HgCl > free CH
3
Hg
+
> CH
3
HgOH; at pH 4.7 to about 7.5, the most common forms
were CH
3
HgCl > CH
3
HgOH > free CH
3
Hg
+
; and at pH 7.5 to 10, this order was CH
3
HgOH >
CH
3
HgCl >> free CH
3
Hg
+
(Reimers and Krenkel, 1974; Jackson, 1998). Reduction of abiotic
inorganic mercury increases with increasing electron donors, low redox potential, and sunlight

intensity (Gabriel and Williamson 2004). Factors that increase mercury volatilization from soils
include increased soil permeability, higher temperatures, and increased sunlight intensity; therefore,
increased volatilization is expected in tropical climates. A decrease in mercury adsorption and an
increase in soil moisture can also increase volatilization (Gabriel and Williamson, 2004).Gabriel
and Williamson (2004) recommend that additional research be conducted on inorganic mer-
cury–ligand formation in water and runoff and its effects on methylmercury formation in soils, and
on quantification of the sources and transport characteristics of methylmercury in terrestrial envi-
ronments. The mercury–ligand form exiting the terrestrial watershed will strongly influence the
mercury/methylmercury bioaccumulation potential in surface waters. Accordingly, more analyses
are needed to determine the mercury forms in terrestrial watershed runoff in dissolved and partic-
ulate fractions (Gabriel and Williamson, 2004).
3.6 MERCURY MEASUREMENT
Techniques for analysis of different mercury species in biological samples and abiotic materials
include atomic absorption, cold vapor atomic fluorescence spectrometry, gas-liquid chromatography
with electron capture detection, neutron activation, and inductively coupled plasma mass spectrom-
etry (Takizawa, 1975; Lansens et al., 1991; Schintu et al., 1992; Porcella et al., 1995). Methyl-
mercury concentrations in marine biological tissues are detected at concentrations as low as 10.0
µg
Hg/kg tissue using graphite furnace sample preparation techniques and atomic absorption spec-
trometry (Schintu et al., 1992).
© 2006 by Taylor & Francis Group, LLC
PROPERTIES 31
3.7 SUMMARY
Physical, chemical, biological, and biochemical properties of mercury are briefly reviewed; mercury
transport and speciation processes are summarized; and analytical techniques listed for mercury
measurement.
In mammals, all forms of mercury can cross the placenta to the fetus and interfere with thiol
metabolism. Chemical speciation is probably the most important variable influencing mercury
toxicity; however, speciation is difficult to quantify. In freshwater lakes, for example, mercury
speciation depends, in part, on pH, alkalinity, redox, and microbial activity. Most authorities agree

that all mercury species discharged into natural bodies of water can be converted into methylmer-
curials — the most toxic form — at rates influenced, in part, by mercury loadings, nutrient content,
sedimentation rates, and suspended sediment loadings. Bioavailability of methylmercury to aquatic
biota is highly dependent on lake chemistry, mercury deposition rates, dissolved organic carbon,
and other variables. Methylmercury, in turn, is decomposed abiotically and by bacteria containing
mercuric reductase and organomercurial lyase; these demethylating strains of bacteria are common
in the environment and have been isolated from water, sediments, soils, and the GI tract of humans
and other mammals. Other mechanisms can reduce inorganic Hg
2+
to Hg
o
in freshwater, with Hg
o
production higher under anoxic conditions, in a chloride-free environment, and at pH 4.5.
Methylmercury concentrations in tissues of marine fishes can now be detected at levels greater
than 10.0
µg/kg tissue using graphite furnace sample preparation techniques and atomic absorption
spectrometry.
REFERENCES
Allard, B. and I. Arsenie. 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.
Anon. 1948. Mercury, Encylop. Brittanica, 15, 269–272.
Baldi, F., F. Semplici, and M. Filippelli. 1991. Environmental applications of mercury resistant bacteria, Water
Air Soil Pollut., 56, 465–475.
Barghigiani, C., D. Pellegrini, and E. Carpene. 1989. Mercury binding proteins in liver and muscle of flat fish
from the northern Tyrrhenian Sea, Comp. Biochem. Physiol., 94C, 309–312.
Barkay, T. 1992. Mercury cycle. In Encyclopedia of Microbiology, Vol. 3, p. 65–74. Academic Press, San Diego.
Becker, D.S. and G.N. Bigham. 1995. Distribution of mercury in the aquatic food web of Onondaga Lake,
New York, Water Air Soil Pollut., 80, 563–571.
Beijer, K. and A. Jernelov. 1979. Methylation of mercury in natural waters. In J.0. Nriagu (Ed.). The Bio-

geochemistry of Mercury in the Environment, p. 201–210. Elsevier/North-Holland Biomedical Press,
New York.
Benoit, J.M., C.C. Gilmour, R.P. Mason, and A. Heyes. 1999a. Sulfide controls on mercury speciation and
bioavailability to methylating bacteria in sediment and pore waters, Environ. Sci. Technol., 33, 951–957.
Benoit, J.M., R.P. Mason, and C.C. Gilmour. 1999b. Estimation of mercury-sulfide speciation and bioavail-
ability in sediment pore waters using octanol-water partitioning, Environ. Toxicol. Chem., 18,
2138–2141.
Berman, M. and R. Bartha. 1986. Levels of chemical versus biological methylation of mercury in sediments,
Bull. Environ. Contam. Toxicol., 36, 401–404.
Birge, W.J., J.A. Black, A.G. Westerman, and J.E. Hudson. 1979. The effect of mercury on reproduction of
fish and amphibians. In J.0. Nriagu (Ed.), The Biogeochemistry of Mercury in the Environment, p.
629–655. Elsevier/North-Holland Biomedical Press, New York.
Bloom, N.S., C.J. Watras, and J.P. Hurley. 1991. Impact of acidification on the methylmercury cycle of remote
seepage lakes, Water Air Soil Pollut., 56, 477–491.
Boudou, A., D. Georgescauld, and J.P. Desmazes. 1983. Ecotoxicological role of the membrane barriers in
transport and bioaccumulation of mercury compounds. In J.0. Nriagu (Ed.), Aquatic Toxicology, p.
117–136. John Wiley, New York.
© 2006 by Taylor & Francis Group, LLC
32 MERCURY HAZARDS TO LIVING ORGANISMS
Boudou, A. and F. Ribeyre. 1983. Contamination of aquatic biocenoses by mercury compounds: an experi-
mental toxicological approach. In J.0. Nriagu (Ed.), Aquatic Toxicology, p. 73–116. John Wiley, New York.
Branfireum, B.A., A. Heyes, and N.T. Roulet. 1996. The hydrology and methylmercury dynamics of a
Precambrian shield headwater peatland, Water Resour. Res., 32, 1785–1794.
Branfireun, B.A., D. Hilbert, and N.T. Roulet. 1998. Sinks and sources of methylmercury in a boreal catchment,
Biogeochemistry, 41, 277–291.
Callister, S.M. and H.R. Winfrey. 1986. Microbial methylation of mercury in upper Wisconsin River sediments,
Water Air Soil Pollut., 29, 453–465.
Choi, S.C., T. Chase, and R. Bartha. 1994. Enzymatic catalysis of mercury methylation by Desulfovibrio
desulfuricans LS, Appl. Environ. Microbiol., 60, 1342–1346.
Clarkson, T.W., R. Hamada, and L. Amin–Zaki. 1984. Mercury. In J.O. Nriagu (Ed.), Changing Metal Cycles

and Human Health, p. 285–309. Springer-Verlag, Berlin.
Clarkson, T.W. and D.O. Marsh. 1982. Mercury toxicity in man. In A.S. Prasad (Ed.), Clinical, Biochemical,
and Nutritional Aspects of Trace Elements, Vol. 6, p. 549–568. Alan R. Liss, Inc., New York.
Compeau, G. and R. Bartha. 1984. Methylation and demethylation of mercury under controlled redox, pH,
and salinity conditions, Appl. Environ. Microbiol., 48, 1202–1207.
Cooper, C.M. and W.B. Gillespie. 2001. Arsenic and mercury concentrations in major landscape components
of an intensively cultivated watershed, Environ. Pollut., 111, 67–74.
Cuvin-Aralar, M.L.A. and R.W. Furness. 1991. Mercury and selenium interaction: a review, Ecotoxicol.
Environ. Safety, 21, 348–364.
Das, S.K., A. Sharma, and G. Talukder. 1982. Effects of mercury on cellular systems in mammals — a review,
Nucleus (Calcutta), 25, 193–230.
Driscoll, C.T., V. Blette, C. Yan, C.L. Schofield, R. Munson, and J. Holsapple. 1995. The role of dissolved
organic carbon in the chemistry and bioavailability of mercury in remote Adirondack lakes, Water Air
Soil Pollut., 80, 499–508.
Ebinghaus, R. H.H. Kock, C. Temme, J.W. Einax, A.G. Lowe, A. Richter, J.P. Burrows, and W.H. Schroeder.
2002. Antarctic springtime depletion of atmospheric mercury, Environ. Sci. Technol., 36, 1238–1244.
Eisler, R. 2000. Mercury. In Handbook of Chemical Risk Assessment: Health Hazards to Humans, Plants,
and Animals, Vol. 1, p. 313–409. Lewis Publishers, Boca Raton, FL.
Eisler, R. 2004. Mercury hazards from gold mining to humans, plants, and animals, Rev. Environ. Contam.
Toxicol., 181, 139–198.
Elhassani, S.B. 1983. The many faces of methylmercury poisoning, J. Toxicol., 19, 875–906.
Falter, R. and R.D. Wilken. 1998. Isotope experiments for the determination of abiotic mercury methylation
potential of a River Rhine sediment, Vom Wasser, 90, 217–232.
Fitzgerald, W.F. 1986. Cycling of mercury between the oceans and the atmosphere. In P. Buat-Menard (Ed.),
The Role of Air-Sea Exchange in Geochemical Cycling, p. 363–408. D. Reidel Publ., Boston.
Fitzgerald, W.F. and T.W. Clarkson. 1991. Mercury and monomethylmercury: present and future concerns,
Environ. Health Perspect., 96, 159–166.
Fitzgerald, W.F., D.R. Engstrom, R.P. Mason, and E.A. Nater. 1998. The case for atmospheric mercury
contamination in remote areas, Environ. Sci. Technol., 32, 1–7.
Gabriel, M.C. and D.G. Williamson. 2004. Principal biogeochemical factors affecting the speciation and

transport of mercury through the terrestrial environment, Environ. Geochem. Health, 26, 421–434.
Gailer, J., G.N. George, I.J. Pickering, S. Madden, R.C. Prince, and E.Y. Yu. 2000. Structural basis of the
antagonism between inorganic mercury and selenium in mammals, Chem. Res. Toxicol., 13, 1135–1142.
Gill, G.A. and K.W. Bruland. 1990. Mercury speciation in surface freshwater systems in California and other
areas, Environ. Sci. Technol., 24, 1392–1400.
Gilmour, C.C., E.A. Henry, and R. Mitchell. 1992. Sulfate stimulation of mercury methylation in freshwater
sediments, Environ. Sci. Technol., 26, 2281–2287.
Gilmour, C.C., G.S. Reidel, M.C. Ederington, J.T. Bell, J.M. Benoit, G.A. Gill, and M.C. Stordal. 1998.
Methylmercury concentrations and production rates across a trophic gradient in the northern Ever-
glades, Biogeochemistry, 40, 327–345.
Goyer, R.A. 1986. Toxic effects of metals. In C.D. Klaassen, M.O. Amdur, and J. Doull (Eds.), Casarett and
Doull’s Toxicology, third edition, p. 582–635. Macmillan, New York.
Greener, Y. and J.A. Kochen. 1983. Methyl mercury toxicity in the chick embryo, Teratology, 28, 23–28.
Grigal, D.F. 2002. Inputs and outputs of mercury from terrestrial watersheds: a review, Environ. Rev., 10, 1–39.
© 2006 by Taylor & Francis Group, LLC
PROPERTIES 33
Hamasaki, T., H. Nagawe, Y. Yoshioka, and T. Sato. 1995. Formation, distribution, and ecotoxicity of meth-
ylmetals of tin, mercury, and arsenic in the environment, Crit. Rev. Environ. Sci. Technol., 25, 45–91.
Hines, M.E., M. Horvat, J. Faganeli, J.C.J. Bonzongo, T. Barkay, E.B. Major, K.J. Scott, E.A. Bailey, J.J
Warwick, and W.B. Lyons. 2000. Mercury biogeochemistry in the Idrija River, Slovenia, from above
the mine into the Gulf of Trieste, Environ. Res., 83A, 129–139.
Hirayama, K. and A. Yasutake. 2001. In vivo degradation of methylmercury: its mechanism and significance
in methylmercury induced neurotoxicity. In Y. Takizawa and M. Osame (Eds.), Understanding of
Minamata Disease: Methylmercury Poisoning in Minamata and Niigata, Japan, p. 103–110. Japan
Public Health Assn., Tokyo.
Hook, J.B. and W.R. Hewitt. 1986. Toxic responses of the kidney. Pages 310–329 In C.D. Klaassen, M.O.
Amdur, and J. Doull (Eds.), Casarett and Doull’s Toxicology, third edition, p. 310–329. Macmillan,
New York.
Huckabee, J.W., J.W. Elwood, and S.G. Hildebrand. 1979. Accumulation of mercury in freshwater biota. In
J.O. Nriagu (Ed.), The Biogeochemistry of Mercury in the Environment. p. 277–302. Elsevier/North-

Holland Biomedical Press, NY.
Hurley, J.P., J.M. Benoit, C.L. Babiarz, M.M. Shafer, A.W. Andren, J. R. Sullivan, R. Hammond, and D. A.
Webb. 1995. Influences of watershed characteristics on mercury levels in Wisconsin rivers, Environ.
Sci. Technol., 29, 1867–1875.
Ikemoto, T., T. Kunito, H. Tanaka, N. Baba, N. Miyazaki, and S. Tanabe. 2004. Detoxification mechanism of
heavy metals in marine mammals and seabirds: interaction of selenium with mercury, silver, zinc,
and cadmium in liver, Arch. Environ. Contam. Toxicol., 47, 402–413.
Itano, K., S. Kawai, N. Miyazaki, R.Tatsukawa, and T. Fujiyama. 1984. Mercury and selenium levels in striped
dolphins caught off the Pacific coast of Japan, Agricul. Biol. Chem., 48, 1109–1116.
Iwata, H., T. Masukawa, H. Kito, and M. Hiyashi. 1982. Degradation of methylmercury by selenium, Life
Sci., 31, 859–866.
Jackson, T.A. 1986. Methyl mercury levels in a polluted prairie river-lake system: seasonal and site-specific
variations, and the dominant influence of trophic conditions, Can. J. Fish. Aquat. Sci., 43, 1873–1887.
Jackson, T.A. 1998. Mercury in aquatic ecosystems. In W.J. Langston and M.J. Bebianno (Eds.), Metal
Metabolism in Aquatic Environments, p. 76–157. Chapman & Hall, London.
Jernelov, A. 1969. Conversion of mercury compounds. In M.W. Miller and G.G. Berg (Eds.), Chemical Fallout,
Current Research on Persistent Pesticides, p. 68–74. Chas. C Thomas, Springfield, IL.
Kim, E.Y., K. Saeki, S. Tanabe, H. Tanaka, and R. Tatsukawa. 1996. Specific accumulation of mercury and
selenium in seabirds, Environ. Pollut., 94, 261–265.
Kim, J.H., P.J. Hanson, M.O. Barnett, and S.E. Lindberg. 1997. Biogeochemistry of mercury in the air-soil-
plant system. In A. Sigel and H. Sigel (Eds.), Metal Ions in Biological Systems, Vol. 34, Mercury and
its Effects on Environment and Biology, p. 185–212. Marcel Dekker, New York.
King, J.K., J.E. Kostka, M.E. Frischer, F.M. Saunders, and R.A. Jahnke. 2001. A quantitative relationship that
demonstrates mercury methylation rates in marine sediments are based on the community composition
and activity of sulfate-reducing bacteria, Environ. Sci. Technol., 35, 2491–2496.
Koeman, J.H, W.H.M. Peeters, C.H.M. Koudstaal-Hol, P.S. Tijoe, and J.J.M. de Goeij. 1973. Mercury-selenium
correlations in marine mammals, Nature, 254, 385–386.
Krabbenhoft, D.P. and C. Babiarz. 1992. The role of groundwater transport in aquatic mercury cycling, Water
Resour. Res., 28, 3119–3128.
Krabbenhoft, D.P., J.M. Benoit, C.L. Babiarz, J.P. Hurley, and A.W. Andrem. 1995. Mercury cycling in the

Allequash Creek watershed, Water Air Soil Pollut., 80, 425–433.
Krabbenhoft, D.P., J.P. Hurley, M.L. Olson, and L.B. Cleckner. 1998. Diel variability of mercury phase and
species distributions in the Florida Everglades, Biogeochemistry, 40, 311–325.
Krabbenhoft, D.P., M.L. Olson, J.F. Dewild, D.W. Clow, R.S. Striegl, M.M. Dornblaster, and P. Van Metre.
2002. Mercury loading and methylmercury production and cycling in high altitude lakes from the
western United States, Water Air Soil Pollut. Focus, 2, 233–249.
Law, R.J. 1996. Metals in marine mammals. In W.N. Beyer, G.H. Heinz, and A.W. Redmon-Norwood (Eds.),
Environmental Contaminants in Wildlife: Interpreting Tissue Concentrations, p. 357–376. CRC Press,
Boca Raton, FL.
Lansens, P., M. Leermakers, and W. Baeyens. 1991. Determination of methylmercury in fish by headspace-
gas chromatography with microwave-induced-plasma detection, Water Air Soil Pollut., 56, 103–115.
© 2006 by Taylor & Francis Group, LLC
34 MERCURY HAZARDS TO LIVING ORGANISMS
Lindberg, S.E., S. Brooks, C.J. Lin, K.J. Scott, M.S. Landis, R.K. Stevens, M. Goodsite, and A. Richter. 2002.
Dynamic oxidation of gaseous mercury in the Arctic troposphere at polar sunrise, Environ. Sci.
Technol., 36, 1245–1256.
Lindberg, S.E. and W.J. Stratton. 1998. Atmospheric mercury speciation: concentrations and behavior of
reactive gaseous mercury in ambient air, Environ. Sci. Technol., 32, 49–57.
Lorey, P. and C.T. Driscoll. 1999. Historical trends of mercury deposition to Adirondack lakes, Environ. Sci.
Technol., 33, 718–722.
Martoja, R. and J.P. Berry. 1980. Identification of tiemannite as a probable product of demethylation of mercury
by selenium in cetaceans: a complement to the scheme of the biological cycle of mercury, Vie Milieu,
30, 7–10.
Marvin-DiPasquale, M., J. Agee, C. McGowan, R.S. Oremland, M. Thomas, D.P. Krabbenhoft, and C.C.
Gilmour. 2000. Methyl mercury degradation pathways — a comparison among three mercury impacted
ecosystems, Environ. Sci. Technol., 34, 4908–4916.
Marvin-DiPasquale, M. and R.S. Oremland. 1998. Bacterial methylmercury degradation in Florida Everglades
sediment and periphyton, Environ. Sci. Technol., 32, 2556–2563.
Mason, R.P. and W.F. Fitzgerald. 1991. Mercury speciation in open ocean waters, Water Air Soil Pollut., 56,
779–789.

Mason, R.P., W.F. Fitzgerald, and F.M.M. Morel. 1994. The biogeochemical cycling of elemental mercury:
anthropogenic influences, Geochim. Cosmochim. Acta, 58, 3191–3198.
Merck Index. 1976. The Merck Index, 9th edition. Merck and Co. Inc., Rahway, NJ.
Munthe, J., I. Wangberg, N. Pirrone, A. Iverfeld, R. Ferrara, R. Ebinghaus, X. Feng, K. Gardfeldt, G. Keeler,
E. Lanzillotta, S. Lindberg, J. Lu, Y. Mamane, E. Prestbo, S. Schmolke, W.H. Schroeder, J. Sommar,
F. Sprovieri, R.K. Stevens, W. Stratton, G. Tuncel, and A. Urba. 2001. Intercomparison of methods
for sampling and analysis of atmospheric mercury species, Atmospher. Environ., 35, 3007–3017.
Nagase, H., Y. Ose, and T. Sato. 1988. Possible methylation of inorganic mercury by silicones in the environ-
ment, Sci. Total Environ., 73, 29–36.
Nakamura, K. 1994. Mercury compounds-decomposing bacteria in Minamata Bay. In Proceedings of the
International Symposium on “Assessment of Environmental Pollution and Health Effects from Meth-
ylmercury,” p. 198–209. October 8–9, 1993, Kumamoto, Japan. National Institute for Minamata
Disease, Minamata City, Kumamoto 867, Japan.
Nigro, M. 1994. Mercury and selenium localization in macrophages of the striped dolphin, Stenella coerule-
oalba, J. Mar. Biol. Assoc. U.K., 74, 975–978.
Nigro, M. and C. Leonzio. 1996. Intracellular storage of mercury and selenium in different marine vertebrates,
Mar. Ecol. Prog. Ser., 135, 137–143.
Olson, B.H. and R.C. Cooper. 1976. Comparison of aerobic and anaerobic methylation of mercuric chloride
by San Francisco Bay sediments, Water Res., 10, 113–116.
Oremland, R.S., C.W. Culbertson, and M.R. Winfrey. 1991. Methylmercury decomposition in sediments and
bacterial cultures involvement of methanogens and sulfate reducers in oxidative demethylation, Appl.
Environ. Microbiol., 57, 130–140.
Pacyna, E.G., J.M. Pacyna, and N. Pirrone. 2001. European emissions of atmospheric mercury from anthro-
pogenic sources in 1995, Atmosph. Environ., 35, 2987–2996.
Pak, K.R. and R. Bartha. 1998. Mercury methylation and demethylation in anoxic lake sediments and by
strictly anaerobic bacteria, Appl. Environ. Microbiol., 64, 1013–1017.
Palmisano, F., N. Cardellicchio, and P.G. Zambonin. 1995. Speciation of mercury in dolphin liver: a two-stage
mechanism for the demethylation accumulation process and role of selenium, Mar. Environ. Res., 40,
109–121.
Pelletier, E. 1985. Mercury-selenium interactions in aquatic organisms: a review, Mar. Environ. Res., 18, 111–132.

Ping, L., H. Nagasawa, K. Matsumoto, A. Suzuki, and K. Fuwa. 1986. Extraction and purification of a new
compound containing selenium and mercury accumulated in dolphin liver, Biol. Trace Elem. Res., 11,
185–199.
Porcella, D., J. Huckabee, and B. Wheatley (Eds.). 1995. Mercury as a global pollutant. Water Air Soil Pollut.,
80, 1–1336.
Ramamoorthy, S. and K. Blumhagen. 1984. Uptake of Zn, Cd, and Hg by fish in the presence of competing
compartments, Can. J. Fish. Aquat. Sci., 41, 750–756.
© 2006 by Taylor & Francis Group, LLC
PROPERTIES 35
Ravichandran, M., G.R. Aiken, J.N. Ryan, and M.M. Reddy. 1999. Inhibition of precipitation and aggregation
of metacinnabar (mercury sulfide) by dissolved organic matter isolated from the Florida Everglades,
Environ. Sci. Technol., 33, 1418–1423.
Rawson, A.J., J.P. Bradley, A. Teetsov, S.B. Rice, E.M. Haller, and G.W Patton. 1995. A role for airborne
particulates in high mercury levels of some cetaceans, Ecotoxicol. Environ. Safety, 30, 309–314.
Regnell, O. 1990. Conversion and partitioning of radio-labelled mercury chloride in aquatic model systems,
Can. J. Fish. Aquat. Sci., 47, 548–553.
Reimers, R.S. and P.A. Krenkel. 1974. Kinetics of mercury adsorption and desorption in sediments, J. Water
Pollut. Control Feder., 46, 352–365.
Rolfhus, K.R. and W.F. Fitzgerald. 1995. Linkages between atmospheric mercury deposition and the meth-
ylmercury content of marine fish, Water Air Soil Pollut., 80, 291–297.
Roulet, M. and M. Lucotte. 1995. Geochemistry of mercury in pristine and flooded ferralitic soils of a tropical
rain forest in French Guiana, South America, Water Air Soil Pollut., 80, 1079–1088.
Rowland, I.R. 1988. Interactions of the gut microflora and the host in toxicology, Toxicol. Pathol., 16, 147–153.
Rowland, I.R., R.D. Robinson, and R.A. Doherty. 1984. Effects of diet on mercury metabolism and excretion
in mice given methylmercury: role of gut flora, Arch. Environ. Health, 39, 401–408.
Schintu, M., F. Jean-Caurant, and J.C. Amiard. 1992. Organomercury determination in biological reference
materials: application to a study on mercury speciation in marine mammals off the Faroe Islands,
Ecotoxicol. Environ. Safety, 24, 95–101.
Schroeder, W.H. and J. Munthe. 1998. Atmospheric mercury — an overview, Atmos. Environ., 32, 809–822.
Schuster, P.F., D.P. Krabbenhoft, D.L. Naftz, L.D. Cecil, M.L. Olson, and J.F. Dewild. 2002. Atmospheric

mercury deposition during the last 270 years: a glacial ice core record of natural and anthropogenic
sources, Environ. Sci. Technol., 36, 2302–2310.
Sellers, P., C.A. Kelly, and J.W.M. Rudd. 2001. Fluxes of methylmercury to the water column of a drainage
lake: the relative importance of internal and external sources, Limnol. Ocean., 46, 623–631.
Sellers, P., C.A. Kelly, J.W.M. Rudd, and A.R. MacHutchon. 1996. Photodegradation of methylmercury in
lakes, Nature, 380, 694–697.
Sellinger, M., N. Ballatori, and J.L. Boyer. 1991. Mechanism of mercurial inhibition of sodium-coupled alanine
uptake in liver plasma membrane vesicles from Raja erinacea, Toxicol. Appl. Pharmacol., 107,
369–376.
Semkin, R.G., G. Mierle, and R.J. Neureuther. 2005. Hydrochemistry and mercury cycling in a high Arctic
watershed, Sci. Total Environ., 342, 199–221.
Sicialiano, S.D., N.J. O’Driscoll, R. Tordon, J. Hill, S. Beauchamp, and D.R.S. Lean. 2005. Abiotic production
of methylmercury by solar radiation, Environ. Sci. Technol., 39, 1071–1077.
Spencer, J.N. and A.F. Voigt. 1961. Thermodynamics of the solution of mercury metal. I. Tracer determination
of the solubility in various liquids, Jour. Phys. Chem., 72, 464–470.
Stratton, W.J., S. Lindberg, and C.J. Perry. 2001. Atmospheric mercury speciation: laboratory and field
evaluation of a mist chamber method for measuring reactive gaseous mercury, Environ. Toxicol. Chem.,
35, 170–177.
Suda, I., S. Totoki, T. Uchida, and H. Takahashi. 1992. Degradation of methyl and ethyl mercury into inorganic
mercury by various phagocytic cells, Arch. Toxicol., 66, 40–44.
Takizawa, Y. 1975. Studies on the distribution of mercury in several body organs — application of activography
to examination of mercury distribution of Minamata Disease patients. In Studies on the Health Effects
of Alkylmercury in Japan, p. 5–9. Environment Agency Japan, Tokyo.
Thompson, D.R. 1990. Metal levels in marine vertebrates. In R.W. Furness and P.S. Rainbow (Eds.), Heavy
Metals in the Marine Environment, p. 143–182. CRC Press, Boca Raton, FL.
U.S. Environmental Protection Agency (USEPA). 1980. Ambient water quality criteria for mercury, U.S.
Environ. Protection Agen. Rep. 440/5-80-058. Available from Natl. Tech. Infor. Serv., 5285 Port Royal
Road, Springfield, VA 22161.
U.S. Environmental Protection Agency (USEPA). 1985. Ambient water quality criteria for mercury — 1984.
U.S. Environ. Protection Agen. Rep. 440/5-84-026. 136 pp. Available from Natl. Tech. Infor. Serv.,

5285 Port Royal Road, Springfield, VA 22161.
U.S. Environmental Protection Agency (USEPA). 1997. Mercury study report to Congress, U.S. Environ.
Protection Agen. Publ. 452R-97 004 Washington, D.C.
© 2006 by Taylor & Francis Group, LLC
36 MERCURY HAZARDS TO LIVING ORGANISMS
U.S. National Academy of Sciences (USNAS). 1978. An Assessment of Mercury in the Environment. Natl.
Acad. Sci., Washington, D.C., 185 pp.
Weast, R.C. and M. J. Astle (Eds.). 1982. CRC Handbook of Chemistry and Physics. CRC Press, Boca Raton,
FL.
Wiener, J.G., D.P. Krabbenhoft, G.H. Heinz, and A.M. Scheuhammer. 2003. Ecotoxicology of mercury. In
D.J. Hoffman, B.A. Rattner, G.A. Burton Jr., and J. Cairns Jr. (Eds.)., Handbook of Ecotoxicology,
2nd edition, p. 409–436. Lewis Publ., Boca Raton, FL.
Windom, H.L. and D.R. Kendall. 1979. Accumulation and biotransformation of mercury in coastal and marine
biota. In J.O. Nriagu (Ed.), The Biogeochemistry of Mercury in the Environment, p. 301–323.
Elsevier/North-Holland Biomedical Press, NY.
Winfrey, M.R. and J.W.M. Rudd. 1990. Environmental factors affecting the formation of methylmercury in
low pH lakes, Environ. Toxicol. Chem., 9, 853-869.
World Health Organization (WHO). 1976. Mercury. Environmental Health Criteria 1. WHO, Geneva, 131 pp.
World Health Organization (WHO). 1989. Mercury — Environmental Aspects. Environmental Health Criteria
86. WHO, Geneva. 115 pp.
World Health Organization (WHO). 1990. Methylmercury. Environmental Health Criteria 101. WHO, Geneva.
144 pp.
World Health Organization (WHO). 1991. Inorganic Mercury. Environmental Health Criteria 118. WHO,
Geneva. 168 pp.
Yang, J., T. Kunito, S. Tanabe, and N. Miyazaki. 2002. Mercury in tissues of Dall’s porpoise (Phocoenoides
dalli) collected on Sanriku coast of Japan, Fish Sci., 68 (Suppl. 1), 256–259.
Yasutake, A. and K. Hirayama. 2001. Evaluation of methylmercury biotransformation using rat liver slices,
Arch. Toxicol., 75, 400–406.
Zhang, H. and S.E. Lindberg. 2001. Sunlight and iron(III)-induced photochemical production of dissolved
gaseous mercury in freshwater, Environ. Sci. Technol., 35, 928–935.

Zillioux, E.J., D.B. Porcella, and J.M. Benoit. 1993. Mercury cycling and effects in freshwater wetland
ecosystems, Environ. Toxicol. Chem., 12, 2245–2264.
© 2006 by Taylor & Francis Group, LLC

×