Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2020 Jan 6.
Published in final edited form as: J Toxicol Environ Health B Crit Rev. 2010;13(5):385–410. doi: 10.1080/10937401003673750

TRANSPORT OF INORGANIC MERCURY AND METHYLMERCURY IN TARGET TISSUES AND ORGANS

Christy C Bridges 1, Rudolfs K Zalups 1
PMCID: PMC6943924  NIHMSID: NIHMS1064960  PMID: 20582853

Abstract

Owing to the prevalence of mercury in the environment, the risk of human exposure to this toxic metal continues to increase. Following exposure to mercury, this metal accumulates in numerous organs, including brain, intestine, kidneys, liver, and placenta. Although a number of mechanisms for the transport of mercuric ions into target organs were proposed in recent years, these mechanisms have not been characterized completely. This review summarizes the current literature related to the transport of inorganic and organic forms of mercury in various tissues and organs. This review identifies known mechanisms of mercury transport and provides information on additional mechanisms that may potentially play a role in the transport of mercuric ions into target cells.


Mercury (Hg) is an extremely toxic group IIB metal found in many occupational and environmental settings. This metal may exist in several different physical and chemical forms. Elemental mercury (Hg0), for instance, exists as a liquid at room temperature, and because of its high vapor pressure, it can be released readily into the atmosphere as Hg vapor. Hg also exists as mercurous (Hg+) or mercuric (Hg2+) salts of anionic species of chlorine, sulfur, or oxygen. Of the inorganic forms of Hg, the mercuric form is the most abundant in environmental settings. Organic species of Hg are formed when a mercuric ion binds covalently with a carbon atom of an organic functional group such as a methyl, ethyl, or phenyl group. Methylmercury (CH3Hg+) is by far the most common form of organic Hg to which humans and animals are exposed. CH3Hg+ in the environment is predominantly formed by methylation of inorganic mercuric ions by microorganisms present in soil and water (Zalups, 2000; Clarkson & Magos, 2006; ATSDR, 2007; Rooney, 2007).

Since Hg is present in the environment in such a ubiquitous manner, it is virtually impossible for humans to avoid exposure to some form of Hg. In addition to environmental exposure, individuals may be exposed to Hg from dental amalgams, medicinal treatments (including vaccinations), and/or dietary sources (Geier & Geier, 2007; Kern & Jones, 2006; Dietert & Dietert, 2008) (Zalups, 2000; Clarkson & Magos, 2006; ATSDR, 2007; Risher & De Rosa, 2007). Of the different routes of exposure, most humans are exposed to Hg following ingestion of food and/or water contaminated with CH3Hg+. High levels of CH3Hg+ are found often in large predatory freshwater and saltwater fish, such as northern pike, salmon, swordfish, tuna, and shark (Bayen et al., 2005; Simmonds et al., 2002). Accumulation of Hg in tissues of these fish is related primarily to the predatory nature and the longevity of these fish in contaminated waters. Fish do not have the same ability to eliminate mercuric species as do mammals. Following ingestion of contaminated food and/or water, CH3Hg+ is absorbed readily by the gastrointestinal tract and enters the systemic circulation, where mercuric ions can be delivered to target organs (ATSDR, 2007).

Many organ systems are affected negatively by exposure to mercuric species. These include, but are not limited to, the cardiovascular (Warkany & Hubbard, 1953; Wakita, 1987; Carmignani et al., 1992; Soni et al., 1992; Wang et al., 2000), gastrointestinal (Lundgren & Swensson, 1949; Afonso & De Alvarez, 1960; Murphy et al., 1979; Bluhm et al., 1992), neurological (Jaffe et al., 1983; Lin & Lim, 1993), hepatobiliary (Murphy et al., 1979; Samuels et al., 1982; Jaffe et al., 1983), and renal (Murphy et al., 1979; Samuels et al., 1982; Rowens et al., 1991) systems.

Our current understanding of the transport of mercuric species is based on several important discoveries. First, many studies showed that the majority of mercuric ions present within biological systems are bound to molecules containing a free sulfhydryl (thiol) group. Mercuric ions rarely exist in a free, unbound state. Second, the cloning and characterization of numerous transport proteins facilitated the study of select transporters and their involvement in the uptake and secretion of mercuric species. Third, technological advances have provided investigators with the ability to study specific transport proteins, by either altering expression levels or isolating an individual transport protein. Each of these discoveries and advances has played an important role in the acquisition of the current body of knowledge, which is summarized in the following review.

BONDING CHARACTERISTICS OF MERCURIC IONS

When considering the handling of mercuric ions in biological systems, one must account for the bonding characteristics of these ions in the various body compartments of humans and other mammals. Mercuric ions have a high affinity for various nucleophilic functional groups, especially the sulfhydryl group that is present in biomolecules such as glutathione (GSH), cysteine (Cys), homocysteine (Hcy), N-acetylcysteine (NAC), and albumin. Due to the predilection of mercuric ions for bonding to the reduced sulfur atom of thiols, one can assume that most mercuric ions within the various tissue and fluid compartments of mammals are bound to sulfhydryl-containing molecules and thus do not exist as inorganic salts, or in an unbound, “free” ionic state (Hughes, 1957).

Inorganic mercuric ions bond with low-molecular-weight thiols in a linear II, coordinate covalent manner (Fuhr & Rabenstein, 1973; Rubino et al., 2004), while organic mercurials, such as CH3Hg+, form linear I, coordinate covalent complexes with these molecules. Thiol S-conjugates of mercuric ions appear to be thermodynamically stable in an aqueous environment possessing a pH ranging from 1 to 14 (Fuhr & Rabenstein, 1973). The affinity constant for mercuric ions bonding to thiolate anions is on the order of 1015 to 1020. Despite the thermodynamic stability of the coordinate covalent bonds formed between mercuric ions and various thiol-containing molecules in aqueous solution, the bonding between mercuric ions and these thiol-containing molecules appears to be more labile within the living organism (Fuhr & Rabenstein, 1973).

Complex factors such as thiol and/or other nucleophilic competition and exchange are likely the causes of the perceived labile nature of bonding that occurs between mercuric ions and certain thiol-containing molecules in particular tissue and cellular compartments. For example, the majority of mercuric ions present in plasma (shortly after exposure to Hg2+) are bound to sulfhydryl-containing proteins, such as albumin (Friedman, 1957; Mussini, 1958; Cember et al., 1968; Lau & Sarkar, 1979). Yet these mercuric ions do not remain bound to these proteins for very long, as evidenced by the rapid decrease in the plasma burden of Hg2+ accompanied by a rapid rate of uptake of mercuric ions in the kidneys, liver, and other organs.

More complex binding arrangements also occur between mercuric ions and protein thiols, such as the metal-containing proteins metallothionein 1 and metallothionein 2 (MT-1 and −2). In each MT-1 or MT-2 molecule, as many as seven atoms of Hg may be bonded coordinately to four sulfur atoms of cysteinyl residues (Zalups, 2000).

Current evidence indicates that mercuric S-conjugates of small endogenous thiols (such as Cys, Hcy, and NAC) are likely the primary transportable forms of mercury in the kidneys. Therefore, it appears that mercuric ions are transferred from plasma proteins to these low-molecular-weight thiols by a complex ligand-exchange mechanism. Moreover, the effectiveness of thiol-containing pharmacological agents, such as penicillamine, N-acetylpenicillamine, meso-2,3-dimercaptosuccinic acid (DMSA), 2,3-dimercapto-1-propanesulfonic acid (DMPS), dithioerythritol, and dithiothreitol, in reversing or protecting against adverse effects of mercury-containing compounds is based on, and can be best explained by, the ability of these agents to remove inorganic and organic mercuric ions from endogenous ligands via nucleophilic competition and exchange, leading to formation of new thiol-mercury complexes.

TRANSPORT OF INORGANIC MERCURY

Intestinal Transport of Hg2+

One route of exposure to Hg2+ is via consumption of food and/or liquids contaminated with Hg2+. Although humans can be exposed frequently to Hg2+ in this manner, the subsequent intestinal absorption is not a major route of absorption. It has been suggested that the means by which Hg2+ is transported across intestinal enterocytes depends upon the species of Hg2+ present in the intestinal lumen, which itself is dependent upon the ligands available to which Hg2+ can bind (Foulkes, 2000). Therefore, it is likely that multiple mechanisms, with varying modes of transport, are involved in the absorption of mercuric ions (Table 1). Ingested food has a high concentration of sulfhydryl-containing molecules, such as amino acids and peptides, which may bond to Hg2+. Thiol S-conjugates of Hg2+ formed within the gastrointestinal tract may act as structural and/or functional homologs of select endogenous molecules, such as amino acids and/or polypeptides that are absorbed along the small intestine. Given the prevalence of amino acid and peptide transporters in enterocytes, especially along the duodenum (Ganapathy et al., 2001; Dave et al., 2004), it is possible that thiol S-conjugates of Hg2+ are transported into cells by one or more of these carriers. Not surprisingly, the initial site of Hg2+ absorption appears to be the duodenum (Endo et al., 1984).

TABLE 1.

Known and Postulated Mechanisms for the Transport of Inorganic Mercury in Various Organs

Organ Direction of transport Known mechanism(s) Postulated mechanism(s) Postulated species of Hg2+
Small intestine Absorption from lumen None at present Amino acid and peptide transporters, DMT1, ZIP8, MRP3 on basolateral membrane S-conjugates of thiol-containing amino acids and peptides, mercuric ions
Secretion into lumen None at present Paracellular transport, amino acid transporters, MRP (e.g., MRP2) S-conjugates of thiol-containing amino acids and/or GSH
Kidney Uptake at apical membrane System b0,+ Other amino acid transporters, endocytosis S-conjugates of Cys, Hcy, albumin and/or other amino acids
Uptake at basolateral membrane OAT1, OAT3 Endocytosis S-conjugates of Cys, Hcy, NAC and/or albumin
Secretion at apical membrane MRP2 MRP4 S-conjugates of thiol-containing amino acids and peptides, as well as DMPS and DMSA
Liver Uptake at sinusoidal membrane None at present Endocytosis, amino acid transporters, GSH transporters S-conjugates of ferritin, albumin, thiol-containing amino acids, and/or GSH
Export at canalicular membrane MRP2 S-conjugates of thiol-containing amino acids and peptides, as well as DMPS and DMSA
Placenta Uptake at apical (maternal) membrane None at present Amino acid transporters S-conjugates of thiol-containing amino acids and peptides and/or GSH
Export at basolateral (fetal) membrane None at present Amino acid transporters S-conjugates of thiol-containing amino acids and peptides and/or GSH
Uptake at basolateral (fetal) membrane None at present OAT4 S-conjugates of thiol-containing amino acids and peptides
Export at apical (maternal) membrane None at present MRP2 S-conjugates of thiol-containing amino acids and peptides, as well as DMPS and DMSA

The transport of Hg2+ across plasma membranes of enterocytes appears to utilize passive and active mechanisms (Andres et al., 2002; Laporte et al., 2002; Hoyle & Handy, 2005). Foulkes and Bergman (1993) suggested that the intestinal absorption of Hg2+ is a two-step process whereby Hg2+ initially binds to the plasma membrane in the form of an anion such as mercuric trichloride (HgCl3). Hg2+ then traverses the plasma membrane and is internalized. Multiple mechanisms, including amino acid, peptide, and drug and ion transporters, may play roles in this uptake. Owing to the abundance of amino acid transporters in the luminal plasma membrane of enterocytes and recent evidence implicating amino acid transporters in the transport of the Cys S-conjugate of Hg2+, Cys-S-Hg-S-Cys, in renal proximal tubular cells (Bridges et al., 2004), one must consider that amino acid and/or peptide transporters may play significant roles in the intestinal absorption of Hg2+. In contrast, MRP3, in the basolateral membrane, was found to transport a number of substrates from the intracellular compartment of enterocytes into the blood (Rost et al., 2002; Prime-Chapman et al., 2004; Shoji et al., 2004; Yokooji et al., 2007; Itagaki et al., 2008; Kitamura et al., 2009). Because of its multispecific nature and its subcellular localization, it may also play a role in the intestinal secretion of thiol S-conjugates of Hg2+.

A small amount of Hg2+ may be taken up following ligand exchange whereby a mercuric ion is removed from its thiol carrier and is taken up by one or more ion transporters. One such transporter is the divalent metal transporter 1 (DMT1). This transporter is localized in the apical membrane of enterocytes (Canonne-Hergaux et al., 1999) and may play a role in the transport of mercuric ions. Although the ability of DMT1 to transport Hg2+ has not been shown directly, studies in mice suggest a role for DMT1 in Hg2+ transport in that decreased expression of intestinal DMT1 corresponded to a fall in the intestinal accumulation of Hg2+ (Ilback et al., 2008).

Intestinal absorption of Hg2+ may also involve a zinc (Zn2+) carrier, such as ZIP8. In vitro studies showed that the activity of this carrier is inhibited by Cd2+ and Hg2+ (Dalton et al., 2005; He et al., 2006; Liu et al., 2008). Although it was found that ZIP8 and other ZIP proteins are present in the intestine, their exact membrane localization has not been determined. Furthermore, their ability to transport Hg2+ has not been shown directly.

The intestine also appears to play an important role in the elimination of Hg2+ in feces, either via secretion of Hg2+ across enterocytes via one or more transport mechanisms (Table 1), or via secretion in bile. It appears that Hg2+ is secreted across enterocytes by paracellular and/or transcellular mechanisms (Sugawara et al., 1998; Zalups et al., 1999a; Hoyle & Handy, 2005). Data from in vivo studies in rats with cannulated or ligated bile ducts indicate that a substantial fraction of the total pool of Hg2+ that is excreted in the feces is due to intestinal secretion of Hg2+ from blood into the intestinal lumen (Zalups, 1998c). This secretion may involve the transport of thiol S-conjugates of Hg2+, which may act as a mimic or a functional homolog of one or more endogenous molecules secreted normally by enterocytes. Potential mechanisms for this secretion include amino acid transporters and multidrug resistance-associated proteins (MRP). Many amino acid transporters are counter-exchangers and thus may mediate bidirectional transport of substrates. In addition, MRP2, which is present in the apical membrane of enterocytes (Maher et al., 2005), was characterized as an export protein, and as such may play a role in the export of Hg2+ from enterocytes into the intestinal lumen. Other members of the MRP family, including MRP4, MRP5, MRP6, and MRP7, have been identified in enterocytes (Maher et al., 2005); however, the membrane localization of these proteins is currently unknown.

Renal Transport of Hg2+

The kidneys are, by far, the primary sites of mercury accumulation following exposure to elemental or inorganic forms of mercury (Adam, 1951; Ashe et al., 1953; Friberg et al., 1957; Friberg, 1959; Berlin & Gibson, 1963; Clarkson & Magos, 1966; Swensson & Ulfvarson, 1968; Cherian & Clarkson, 1976; Zalups & Diamond, 1987a, 1987b; Hahn et al., 1989, 1990; Zalups & Barfuss, 1990; Zalups, 1991a, 1991b, 1991c, 1993a). Renal uptake and accumulation of mercury in vivo occurs rapidly with as much as 50% of a low (0.5 μmol kg−1) dose of Hg2+ present in kidneys of rats within a few hours after exposure (Zalups, 1993a).

Within the kidneys, Hg2+ accumulates primarily in the cortex and outer stripe of the outer medulla (Berlin & Ullberg, 1963c, 1963b, 1963a; Taugner, 1966; Zalups & Diamond, 1987a; 1987b; Zalups & Barfuss, 1990; Zalups, 1991a, 1991c, 1993a). Histochemical and autoradiographic data from mice and rats (Taugner et al., 1966; Hultman et al., 1985, 1992; Magos et al., 1985; Hultman & Enestrom, 1986, 1992; Rodier et al., 1988; Zalups, 1991a) and tubular microdissection data from rats and rabbits (Zalups & Barfuss, 1990; Zalups, 1991b) indicate that the accumulation of Hg2+ in the renal cortex and outer stripe of the outer medulla occurs almost exclusively along the convoluted and straight segments of the proximal tubule. Deposits of mercury have also been localized in the renal proximal tubule of monkeys exposed to elemental mercury from dental amalgams (Danscher et al., 1990). Although the proximal tubule is the primary site where mercuric ions are taken up and accumulated, there are insufficient data to exclude the possibility that other renal tubular segments may also take up, accumulate, and secrete mercuric ions.

A series of in vitro studies provides definitive evidence related to the mechanisms involved in the proximal tubular uptake of mercury. Data from these studies indicate that there are luminal and basolateral mechanisms involved in the uptake of mercuric ions by proximal tubular epithelial cells (Table 1) (Zalups & Barfuss, 1993, 1995, 1998a, 1998b; Zalups, 1995, 1997, 1998a, 1998b; Zalups & Minor, 1995; Zalups & Lash, 1997a; Wei et al., 1999; Aslamkhan et al., 2003; Bridges et al., 2004; Bridges & Zalups, 2004; Zalups & Ahmad, 2005a, 2005b, 2005c). Luminal uptake of Hg2+ appears to be strongly dependent upon the actions of γ-glutamyltransferase and cysteinylglycinase. Indeed, inhibition of γ-glutamyltransferase reduces significantly the renal uptake and accumulation of Hg2+ (Berndt et al., 1985; Tanaka et al., 1990; Tanaka-Kagawa et al., 1993; de Ceaurriz et al., 1994; Zalups, 1995). It was postulated that some GSH S-conjugates of Hg2+ (G-S-Hg-S-G) are filtered at the glomerulus and delivered into the lumen of proximal tubules, where they are degraded rapidly and sequentially by γ-glutamyltransferase and cysteinylglycinase to yield Cys S-conjugates of Hg2+ (Cys-S-Hg-S-Cys).

In addition to the findings referenced earlier, there are several sets of data indicating that the intracellular content of GSH (and likely other thiols) exerts a significant influence on the uptake and accumulation of Hg2+ in renal tubular epithelial cells and hepatocytes. In the kidney, chemically induced depletion of intracellular GSH with either buthionine sulfoximine (which inhibits the γ-glutamylcysteine synthetase) or diethyl maleate (which binds GSH) was shown to decrease the accumulation of Hg2+ in tubular epithelial cells in both the renal cortex and outer stripe of the outer medulla (Berndt et al., 1985; Baggett & Berndt, 1986; Zalups & Lash, 1997b; Zalups et al., 1999a, 1999b). Moreover, studies in right-side-out, brush-border membrane vesicles (isolated from the renal cortex and outer stripe of the outer medulla of rats) indicate that mercuric ions are taken up more readily when they are bound to Cys than when they are conjugated to GSH or the dithiol chelator, DMPS (Zalups & Lash, 1997a). Studies using isolated perfused proximal tubules from rabbits (Cannon et al., 2000) provided additional evidence for the luminal uptake of Cys-S-Hg-S-Cys.

Subsequent experiments in isolated, perfused proximal tubules from rabbits showed that luminal uptake of Cys-S-Hg-S-Cys involves at least one Na+-dependent and one Na+-independent amino acid carrier (Cannon et al., 2001). Since Cys-S-Hg-S-Cys and the amino acid cystine are similar in size and shape, it was hypothesized that Cys-S-Hg-S-Cys may act as a molecular mimic of cystine at the site of one or more cystine transporters located in the luminal plasma membrane of proximal tubular epithelial cells. A likely candidate for the Na+-independent transport of Cys-S-Hg-S-Cys is system b0,+. This heterodimeric transporter is comprised of two subunits, b0,+AT and rBAT, and has a high affinity for cystine and neutral and basic amino acids (Palacin et al., 1998; 2001). Recent studies utilizing type II Madin–Darby canine kidney (MDCKII) cells transfected stably with each subunit of system b0,+ indicate that Cys-S-Hg-S-Cys and Hcy-S-Hg-S-Hcy are transportable substrates of this carrier (Bridges et al., 2004; Bridges & Zalups, 2004). In contrast, it appears that mercuric conjugates of GSH (G-S-Hg-S-G), N-acetylcysteine (NAC-S-Hg-S-NAC), and cysteinylglycine (CysGly; CysGly-S-Hg-S-CysGly) are not transported readily by system b0,+ (Bridges et al., 2004). Together, these data provide strong evidence supporting the hypothesis that Cys-S-Hg-S-Cys and Hcy-S-Hg-S-Hcy act as molecular mimics or homologs of the amino acids cystine and homocystine respectively, at the site of system b0,+.

Transport of Hg2+ from peritubular blood into the intracellular compartment of proximal tubular cells accounts for approximately 40–60% of proximal tubular uptake of Hg2+ (Zalups, 1995, 1998a, 1998b; Zalups & Barfuss, 1995, 1998a, 1998b). With in vivo experiments in rats where glomerular filtration, and hence luminal uptake, was reduced to negligible levels, the renal tubular uptake of Hg2+ decreased by 40% (Zalups & Minor, 1995). Based on this finding, it was suggested that the remaining uptake of Hg2+ occurs at the basolateral membrane (Zalups & Minor, 1995). These studies also show that when animals are treated with the organic anion para-aminohippurate (PAH), which specifically inhibits members of the organic anion transporter (OAT) family (Shimomura et al., 1981; Ferrier et al., 1983; Ullrich et al., 1987; Pritchard, 1988), the uptake of Hg2+ is inhibited significantly. Therefore, it is likely that one or more members of the OAT family mediate a significant portion of the basolateral uptake of Hg2+ (Table 1). Numerous in vivo and in vitro studies provide experimental evidence indicating that mercuric conjugates of Cys, Hcy, and NAC are taken up by OAT1 and/or OAT3, which have both been localized in the basolateral plasma membrane of proximal tubular epithelial cells (Kojima et al., 2002; Motohashi et al., 2002). OAT1 appears to be the major mechanism involved in the basolateral uptake of Hg2+ into proximal tubular cells (Zalups & Lash, 1994; Zalups & Barfuss, 1995, 1998a, 1998b; Zalups, 1995, 1998a, 1998b). Indeed, recent findings from MDCK II cells transfected stably with OAT1 showed that Cys-S-Hg-S-Cys (Aslamkhan et al., 2003; Zalups et al., 2004), NAC-S-Hg-S-NAC (Zalups & Barfuss, 1990), and Hcy-S-Hg-S-Hcy (Zalups & Ahmad, 2004) are transportable substrates of this carrier. Additional experiments using Xenopus laevis oocytes implicate OAT3 in the uptake of Cys-S-Hg-S-Cys (Aslamkhan et al., 2003; Zalups et al., 2004). Taken together, the aforementioned data provide strong support for the theory that OAT1 and OAT3 play significant roles in the basolateral uptake of select mercuric species.

A substantial body of evidence has shown that mercuric ions are extracted effectively from renal tubular cells by the dithiol, metal chelators, DMPS and DMSA (Planas-Bohne, 1981; Aposhian, 1983; Aposhian et al., 1992; Zalups, 1993b; Bridges et al., 2008a, 2008b; Ruprecht, 2008; Zalups & Bridges, 2009). Although this extraction appears to involve a direct secretory process whereby mercuric ions move directly from blood into the tubular lumen (Diamond et al., 1988), the exact mechanisms involved in this process have been identified only recently (Table 1). Initial studies showed that the mechanism or mechanisms responsible for the secretion of Hg2+ into the lumen of proximal tubular cells required that Hg2+ be co-transported with GSH (Tanaka-Kagawa et al., 1993). One possible mechanism for this transport is MRP2. This carrier is localized in the luminal plasma membrane of proximal tubular cells (Schaub et al., 1997; 1999) and appears to require GSH for the transport of some substrates (Leslie et al., 2005). Indeed, indirect evidence from Eisai hyperbilirubinemic rats, which lack functional MRP2, suggests that MRP2 plays a role in the hepatobiliary secretion of Hg2+ (Sugawara et al., 1998). Therefore, MRP2 may play a similar role in proximal tubular epithelial cells. A study utilizing proximal tubules from killifish found that the expression of MRP2 increases after exposure to HgCl2 (Terlouw et al., 2002), suggesting that mercuric ions somehow enhance transcription of the mrp2 gene. The ability of MRP2 to transport Hg2+, however, was not measured directly. In addition, in this study proximal tubules were exposed to Hg2+ as HgCl2, which is not a physiologically relevant form of Hg2+. Under normal in vivo conditions, Hg2+ is usually bound to one or more thiol-containing molecules. A more recent study utilized MRP2-deficent (TR) rats to study the role of MRP2 in the secretion of mercuric ions. Rats were exposed initially to HgCl2 and, in order to facilitate the secretion of mercuric ions, were treated subsequently with DMPS or DMSA. These studies demonstrated that MRP2 plays a significant role in the secretion of mercuric ions and the following model for the DMPS- and DMSA-mediated extraction of mercuric ions was proposed (Bridges et al., 2008a, 2008b). DMPS and DMSA are thought to be taken up at the basolateral membrane of proximal tubular cells via OAT1 and the sodium-dependent dicarboxylate transporter (NaC2), respectively (Islinger et al., 2001; Bahn et al., 2002; Burckhardt et al., 2002). Once internalized, DMPS and DMSA likely form complexes with intracellular Hg2+; these complexes appear to be exported by MRP2 across the luminal plasma membrane of proximal tubular cells into the tubular lumen. Indeed, studies utilizing membrane vesicles prepared from Sf9 cells transfected with human MRP2 provided direct evidence that MRP2 is able to transport both DMPS- and DMSA-S-conjugates of Hg2+ (Bridges et al., 2008a, 2008b). Although data from studies in other organs and experimental systems suggest that secretion of mercuric ions via MRP2 requires co-transport of GSH, current renal data do not indicate that co-transport of GSH is required for export of mercuric ions from proximal tubular cells.

Hepatic Transport of Hg2+

Shortly after exposure to Hg2+, mercuric ions in sinusoidal blood are likely bound mainly to plasma proteins (like albumin, ferritin, and γ-globulins) and, to a lesser degree, several nonprotein thiols. There are a few potential mechanisms that can explain the uptake of Hg2+ across the sinusoidal membrane of hepatocytes (Table 1). One of these mechanisms is endocytosis. Fluid-phase, absorptive and receptor-mediated endocytotic processes are involved in the uptake of extracellular molecules (present in sinusoidal blood) into hepatocytes. Significant amounts of fluid uptake and membrane turnover in hepatocytes occur by fluid-phase and receptor-mediated endocytosis (Oka et al., 1989). As is the case with iron (Fe), mercuric ions may gain entry into hepatocytes via endocytosis mediated by one of the Fe-binding proteins and/or albumin. As a matter of fact, ferritin appears to bind Cd, Zn, Be, and Al. Thus, endocytosis of Hg-ferritin or Hg-albumin complexes may indeed serve as a route of entry of Hg2+ into hepatocytes. It has actually been suggested that ferritin may serve a detoxifying protein due its ability to bind a number of cationic forms of several elements (Joshi et al., 1989).

Since Hg2+ forms thermodynamically stable complexes with GSH and/or amino acids, such as Cys and Hcy, these complexes may enter hepatocytes via transporters that mediate uptake of structurally similar endogenous compounds. Within the sinusoidal membrane of hepatocytes, several members of the MRP family that transport conjugates of GSH have been identified (Ballatori et al., 2005; Deeley et al., 2006). However, transport of Hg2+ by these isoforms has not been studied. Furthermore, since these carriers have been characterized as export proteins, it is not clear whether they are capable of mediating uptake of substrates from blood into hepatocytes. In addition to GSH transporters, numerous amino acid carriers have been identified in the liver (Bode, 2001; Wagner et al., 2001), yet it is currently unclear which, if any, of these carriers are present in the sinusoidal membrane.

Unlike transport at the sinusoidal plasma membrane, transport of Hg2+ across the canalicular plasma membrane of hepatocytes has been studied quite extensively. Early studies suggest that hepatobiliary transport of Hg2+ is dependent upon hepatocellular concentrations of GSH (Ballatori & Clarkson, 1983, 1984, 1985a, 1985b; Dutczak & Ballatori, 1992; Zalups & Lash, 1997b). Indeed, when GSH levels in rats were decreased artificially, hepatocellular retention and/or uptake of intravenously administered Hg2+ increased significantly (Zalups & Lash, 1997). These findings suggest that when cytosolic GSH is low, mercuric ions are unable to exit hepatocytes efficiently. Consequently, mercuric ions accumulate in intracellular compartments of hepatocytes. These data also indicate that intracellular levels of GSH play an important role in the transport of Hg2+ out of hepatocytes. The exact mechanisms involved in the hepatobiliary transport of mercuric ions have not been identified; however, experimental evidence suggests that Hg2+, as a conjugate of GSH, is exported across the canalicular membrane. Since G-S-Hg-S-G is similar structurally to GSSG, it may be taken up by a GSSG transporter. MRP2 appears to transport GSSG and is present in the canalicular membrane of hepatocytes (Akerboom et al., 1991; Buchler et al., 1996; Keppler & Konig, 1997). Recently, MRP2 has been implicated in the transport of Hg2+ as a conjugate of DMPS or DMSA (Bridges et al., 2008a, 2008b). Therefore, it is likely that this carrier may also mediate the transport of other species of Hg2+ (Table 1).

Transport of Hg2+ in the Placenta

Maternal exposure to organic forms of Hg is clearly more toxic to developing fetuses than exposure to Hg2+. However, mercuric ions accumulate in placentas of pregnant women exposed to Hg2+ and may be detrimental to the fetus (Inouye & Kajiwara, 1990; Ask et al., 2002). Surprisingly, the mechanisms by which Hg2+ is transported across the placenta are not well defined. Experiments in brush-border membrane vesicles from human placenta suggest that an amino acid transporter may be involved in the placental transport of Hg2+ (Iioka et al., 1987). However, since these studies used HgCl2 instead of a more physiologically relevant species of Hg2+, such as Cys-S-Hg-S-Cys, these findings may not truly represent processes occurring in vivo. Even so, considering these data and the prevalence of amino acid transporters in the placenta (Jansson, 2001; Kudo & Boyd, 2002), it is postulated that Hg2+, when bound to a Cys or Hcy, may be transported across apical and basolateral membranes of placental epithelial cells via one or more amino acid transporters (Table 1). Other transporters, such as OAT4, may mediate the transport of mercuric ions from the fetal circulation into placental trophoblasts; however, the ability of OAT4 to mediate the transport of mercuric species has not been examined. On the apical plasma membrane, MRP2 may mediate the transport of mercuric species from within placental trophoblasts into maternal circulation (Table 1).

TRANSPORT OF METHYLMERCURY

Transport of CH3Hg+ in the Brain

Clinically relevant adverse effects of CH3Hg+ occur in the brain and central nervous system (CNS) (WHO, 2000; ATSDR, 2007). Therefore, it is not surprising that a large number of studies have focused on mechanisms by which CH3Hg+ gains access to the CNS, specifically, how CH3Hg+ crosses the blood–brain barrier. Similar to Hg2+, CH3Hg+ does not exist as a free, unbound cation in biological systems (Hughes, 1957), but rather is found conjugated to thiol-containing biomolecules, such as GSH, Cys, Hcy, NAC, or albumin (Clarkson, 1993). Early studies utilizing homogenates of rat cerebrum demonstrated that GSH is the primary nonprotein thiol bound to CH3Hg+ (Thomas & Smith, 1979). Subsequent studies in rats and primary cultures of bovine brain endothelial cells showed that co-administration of Cys with CH3Hg+ increased the uptake of CH3Hg+ into capillary endothelial cells of the blood–brain barrier (Hirayama, 1980; Aschner & Clarkson, 1988, 1989). Interestingly, the uptake of CH3Hg+ into cells was inhibited significantly by the presence of the neutral amino acid phenylalanine (Hirayama, 1975, 1980, 1985; Thomas & Smith, 1982). In vivo studies in rat brain (Aschner & Clarkson, 1988) and in vitro studies in bovine cerebral capillary endothelial cells (Aschner & Clarkson, 1989) demonstrated that neutral amino acids are capable of inhibiting the uptake of CH3Hg-S-Cys. These data collectively led to the hypothesis that Cys S-conjugates of CH3Hg+ (CH3Hg-S-Cys) are transportable substrates of a neutral amino acid transporter in the capillary endothelium of the blood–brain barrier. It was suggested that the structural similarities between CH3Hg-S-Cys and methionine (Landner, 1971; Jernelov, 1973) allow this species of CH3Hg+ to cross the blood–brain barrier. One possible mechanism for this transport is the amino acid carrier, system L (Table 2) (Wagner et al., 2001).

TABLE 2.

Known and Postulated Mechanisms for the Transport of Methylmercury in Various Organs and Cells

Organ Direction of transport Known mechanism(s) Postulated mechanism(s) Postulated species of CH3Hg+
Brain Uptake at blood-brain barrier System L Other amino acid transporters S-conjugates of thiol-containing amino acids
Erythrocytes Uptake None at present OATs, D-glucose diffusive transporter, Cys facilitated transporter, Cl transporter S-conjugates of GSH and/or Cys
Intestine Absorption from lumen None at present OATs, amino acid and peptide transporters; MRP3 on basolateral membrane S-conjugates of GSH, Cys, and/or CysGly
Secretion into lumen None at present Amino acid transporters (e.g. System L), GSH transporters S-conjugates of Cys and/or GSH
Kidney Uptake at apical membrane System B0,+ Other amino acid transporters S-conjugates of Cys and Hcy
Uptake at basolateral membrane OAT1 OAT3 S-conjugates of Cys, Hcy, and NAC
Secretion at apical membrane MRP2 MRP4 S-conjugates of thiol-containing amino acids, peptides and/or DMPS and DMSA
Liver Uptake at sinusoidal membrane None at present Amino acid transporters, GSH transporters S-conjugates of thiol-containing amino acids, and/or GSH
Export at canalicular membrane MRP2 S-conjugates of thiol-containing amino acids, peptides and/or DMPS and DMSA
Placenta Uptake at apical (maternal) membrane None at present Amino acid transporters S-conjugates of thiol-containing amino acids
Export at basolateral (fetal) membrane None at present Amino acid transporters S-conjugates of thiol-containing amino acids
Uptake at basolateral (fetal) membrane None at present OAT4 S-conjugates of thiol-containing amino acids and GSH
Export at apical (maternal) membrane None at present MRP2 S-conjugates of thiol-containing amino acids, peptides, and/or DMPS and DMSA

System L is a heterodimeric protein, comprised of a heavy chain, 4F2hc, and a light chain, LAT1 or LAT2, bound together by a disulfide bond (Chillaron et al., 2001; Wagner et al., 2001). LAT1 and LAT2 have been localized in the apical and basolateral plasma membranes, respectively, of endothelial cells lining the blood–brain barrier (Betz & Goldstein, 1978). System L is capable of transporting a broad range of substrates (Oldendorf, 1973). Thus, it is possible that this carrier may also utilize CH3Hg-S-Cys as a substrate. Indeed, in vivo studies in rats (Kerper et al., 1992) and in vitro studies utilizing primary cultures of rat astrocytes (Aschner et al., 1990; 1991) suggest that CH3Hg-S-Cys is a transportable substrate of system L. Subsequent studies showed that Hcy S-conjugates of CH3Hg+ (CH3Hg-S-Hcy) may also be substrates of system L (Mokrzan et al., 1995). Specific studies of LAT1 and LAT2 in Xenopus laevis oocytes provide the first direct molecular evidence showing that CH3Hg-S-Cys is a transportable substrate of LAT 1 and 2 (Simmons-Willis et al., 2002). These data also provide substantive evidence for the phenomenon of molecular mimicry in that CH3Hg-S-Cys appears to mimic methionine at the site of system L.

Transport of CH3Hg+ in Erythrocytes

The handling of mercuric ions, specifically GSH S-conjugates of CH3Hg+ (CH3Hg-S-G), has also been studied in erythrocytes. Wu (1995) reported that when experiments were performed at 5°C, multiple transport systems appeared to be involved in the uptake of CH3Hg-S-G. The primary mechanism for this transport was postulated to be a member of the OAT family. Additional mechanisms may include a D-glucose diffusive transporter, a cysteine-facilitated transporter, and/or a Cl transporter (Table 2) (Wu, 1995). Because these experiments were carried out at 5°C, it is possible that the mechanisms for CH3Hg-S-G transport that were identified in the aforementioned study may differ from those responsible for this uptake at physiological temperatures. In subsequent studies, the uptake of CH3Hg-S-G was measured at 5°C and 20°C, and it was suggested that a member of the OAT family was the primary mechanism responsible for CH3Hg-S-G uptake at both temperatures (Wu, 1996). Subsequent studies showed that the uptake of CH3Hg-S-G was inhibited by probenecid, which suggested that this conjugate is a transportable substrate of OAT (Wu, 1997). It should be noted, however, that probenecid is not a specific inhibitor of OAT. Wu (1997) reported that system N, system y+, and the oligopeptide H+ transport system were not involved in the uptake of CH3Hg-S-G.

Intestinal Transport of CH3Hg+

As mentioned previously, humans are exposed to CH3Hg+ primarily through the ingestion of contaminated food and/or water. Unlike the intestinal absorption of Hg2+, absorption of CH3Hg+ by intestinal enterocytes is more efficient. Therefore, it is important that the mechanisms involved in the intestinal absorption of CH3Hg+ are understood thoroughly. A number of various mechanisms may be involved in this process (Table 2). Urano and colleagues (1990) suggested that there are two independent transport systems for the uptake of CH3Hg+, when it is presented to enterocytes as CH3Hg-S-G. Since the luminal uptake of CH3Hg+ can be inhibited by acivicin (an alkylator of γ-glutamyltransferase) and probenecid (a known inhibitor of OAT), it was suggested that one mode of CH3Hg+ uptake into enterocytes is dependent upon the activity of γ-glutamyltransferase, while the other appears to involve one or more members of the OAT family. As noted earlier, probenecid is not a specific inhibitor of OAT, and thus, may inhibit other transporters. Currently, only OAT2 and OAT10 have been detected in the intestine (Hilgendorf et al., 2007; Bahn et al., 2008); the membrane localization of either protein has not been examined. In addition, several members of the OATP family (Hilgendorf et al., 2007) and a novel organic anion transporter-like protein (OATLP1) are present in the intestine (Jung et al., 2006). Currently, there are no published data regarding the ability of these carriers to transport mercuric ions.

Despite the apparent ability of CH3Hg-S-G to be taken up by enterocytes, it appears that CH3Hg+ is absorbed more readily when it is conjugated to one of the products of GSH catabolism (i.e., CysGly or Cys) (Urano et al., 1990). Indeed, when the activity of γ-glutamyltransferase was inhibited, the transport of CH3Hg-S-G into enterocytes was reduced by 50%. The CysGly-S-conjugate of CH3Hg+ (CH3Hg-S-CysGly), which is the first product following the action of γ-glutamyltransferase on CH3Hg-S-G, is likely degraded further at the luminal plasma membrane of enterocytes to yield CH3Hg-S-Cys. Any CH3Hg-S-CysGly that escapes degradation, however, may be taken up into cells via a peptide transporter present in the luminal plasma membrane. In the intestine, di- and tripeptide transporters are the primary means for the uptake of amino acids. Considering this, and the structural resemblance of CH3Hg-S-CysGly to a small peptide, it is possible that this complex mimics an endogenous di- or tripeptide in order to cross the luminal membrane of enterocytes. In contrast, CH3Hg-S-Cys, which appears to be the primary species of CH3Hg+ secreted into the intestine from bile, is absorbed rapidly by enterocytes (Norseth & Clarkson, 1971), possibly by one of the many amino acid carriers present in the luminal membrane of enterocytes.

Clearly, CH3Hg+ is transported out of enterocytes at the basolateral membrane and subsequently enters the circulation (Leaner & Mason, 2002). Unfortunately, the mechanisms involved in this transport remain unclear. It was suggested that the basolateral efflux of CH3Hg+ involves one or more active-transport carrier proteins. Indeed, competitive inhibition experiments in isolated, perfused catfish intestines provide indirect evidence suggesting a role for a neutral amino acid transporter in the basolateral flux of CH3Hg-S-Cys from intestinal enterocytes. Since system L was implicated in the transport of CH3Hg-S-Cys in other cell types and organs, and because of its basolateral localization in enterocytes (Dave et al., 2004), it is a likely candidate for the export of mercuric species.

Alternatively, Foulkes (1993) suggested that intracellular concentrations of GSH play an important role in regulating the basolateral efflux of CH3Hg+ from enterocytes into the circulation. Owing to structural similarities between CH3Hg-S-G and GSH, it may be postulated that a GSH transporter, such as MRP3, may play a role in the basolateral efflux of CH3Hg-S-G.

Renal Transport of CH3Hg+

CH3Hg+ is capable of inducing significant detrimental effects in the kidney (Prickett et al., 1950; Friberg, 1959; Norseth & Clarkson, 1970a, 1970b; Magos & Butler, 1976; Magos et al., 1981, 1985; McNeil et al., 1988), even though the level of accumulation, following acute exposures, is much less than the level that occurs after exposure to inorganic or elemental forms of mercury. Until recently, the mechanism by which CH3Hg+ is taken up by renal tubular epithelial cells remained unknown. Early studies showed that the renal tubular uptake of CH3Hg+ is dependent upon intracellular concentrations of GSH (Richardson & Murphy, 1975). Several additional studies demonstrated that the renal uptake and accumulation of CH3Hg+ increases following co-administration of CH3Hg+ and GSH (Alexander & Aaseth, 1982; Tanaka et al., 1992). It has also been suggested that γ-glutamyltransferase and cysteinylglycinase, which are both present in the luminal membrane of proximal tubular cells, act upon CH3Hg-S-G to yield CH3Hg-S-Cys (Zalups, 2000), which is likely the most transportable form of CH3Hg+. It is important to note that during the catabolism of GSH, the methylmercuric ion remains bound to the sulfur atom of Cys (Naganuma et al., 1988). When the activity of γ-glutamyltransferase was inhibited by acivicin, the uptake of CH3Hg+ into renal tubules decreased while the urinary excretion of GSH and CH3Hg+ increased (Berndt et al., 1985; Mulder & Kostyniak, 1985; Gregus et al., 1987; Naganuma et al., 1988; Yasutake et al., 1989; de Ceaurriz & Ban, 1990; Di Simplicio et al., 1990; Tanaka et al., 1990, 1991, 1992). In addition, in vivo studies in mice deficient in γ-glutamyltransferase showed that less CH3Hg+ was absorbed by renal tubular cells (Ballatori et al., 1998). These data indicate that the catabolism of CH3Hg-S-G is required for the luminal absorption of CH3Hg+ by proximal tubular cells and support the theory that CH3Hg-S-Cys is the most readily transportable species of CH3Hg+.

CH3Hg-S-Cys and methionine are similar structurally; therefore, it is possible that both compounds are transported by the same carrier. Thus, it is not surprising to find that the amino acid transporter, system B0,+ is capable of mediating the transport of CH3Hg+ as a conjugate of Cys or Hcy (CH3Hg-S-Hcy; Table 2) (Bridges & Zalups, 2006). System B0,+ is localized in the luminal plasma membrane of proximal tubular cells (Gonska et al., 2000) and mediates the Na+-dependent transport of many neutral and cationic amino acids, including Met (Sloan & Mager, 1999; Nakanishi et al., 2001). Studies using Xenopus laevis oocytes indicate that system B0,+ is capable of transporting CH3Hg-S-Cys and CH3Hg-S-Hcy in a concentration- and time-dependent manner (Bridges & Zalups, 2006). Uptake of each mercuric species was inhibited by known substrates for system B0,+. It is interesting to note that the substrate specificity of system B0,+ is similar to that of system b0,+, which was found to transport conjugates of Hg2+ (Bridges et al., 2004; Bridges & Zalups, 2004).

In addition to mechanisms on the luminal plasma membrane, there are also basolateral mechanisms involved the renal tubular uptake of CH3Hg+ (Tanaka et al., 1992). Basolateral transport of CH3Hg+ from the peritubular capillaries into proximal tubular cells is thought to involve a multispecific carrier, such as the organic anion transporter 1 (OAT1; Table 2). OAT1 is localized exclusively in the basolateral membrane of proximal tubular epithelial cells (Kojima et al., 2002; Motohashi et al., 2002) and mediates the uptake of Cys-, Hcy-, NAC- and DMPS-S-conjugates of CH3Hg+ (Koh et al., 2002; Zalups & Ahmad, 2005a, 2005b, 2005c).

Interestingly, several studies showed that a significant fraction of Hg in the kidneys of animals exposed to methylmercury is in the inorganic form (Gage, 1964; Norseth & Clarkson, 1970a, 1970b; Omata et al., 1980; Magos et al., 1985; Rodier et al., 1988). These findings suggest that organic mercury is oxidized to inorganic mercury prior to and/or after it enters the renal tubular epithelial cells. In addition, some evidence suggests that methylmercury is converted intracellularly to Hg2+ (Dunn & Clarkson, 1980). The mechanism responsible for this conversion, however, is unknown currently.

It is well documented that the metal chelators DMPS and DMSA are capable of extracting mercuric ions following exposure to CH3Hg+ (Aposhian, 1983; Aposhian et al., 1992). The mechanisms by which this extraction occurs have been unclear until recently. In vivo studies in MRP2-deficient (TR) rats exposed to CH3Hg+ and treated subsequently with NAC, DMPS, or DMSA provide support for the hypothesis that MRP2 mediates the transport of CH3Hg+ from within proximal tubular cells into the tubular lumen (Table 2) (Madejczyk et al., 2007; Zalups & Bridges, 2009). Experiments using membrane vesicles isolated from kidneys of TR rats provided more direct evidence suggesting that CH3Hg-S-NAC is a transportable substrate of MRp2 (Madejczyk et al., 2007). Additional experiments using inside-out membrane vesicles from Sf9 cells transfected with human MRP2 also provide direct evidence indicating that DMPS- and DMSA-S-conjugates of CH3Hg+ are transportable substrates of MRP2 (Zalups & Bridges, 2009). Collectively, these data provide strong support for the hypothesis that MRP2 plays a significant role in the renal elimination of mercuric ions following exposure to CH3Hg+ with subsequent chelation therapy.

Hepatic Transport of CH3Hg+

Following absorption by the intestine, CH3Hg+ is delivered to the liver via portal blood. Little is known about the mechanisms by which CH3Hg+ is transported into hepatocytes at the sinusoidal membrane. In vivo studies in rats showed that hepatic uptake and accumulation of CH3Hg+ was enhanced when Cys or GSH was either co-administered with or subsequently administered to CH3Hg+ (Thomas & Smith, 1982), suggesting that amino acid carriers and/or a GSH transporter may be involved in this process (Table 2). More recently, an in vitro study using a human hepatic cell line (HepG2 cells) demonstrated that cellular uptake of CH3Hg+ occurs more rapidly when cells are exposed to CH3Hg+ as a conjugate of Cys (Wang et al., 2000). The specific mechanisms involved in this transport remain unknown.

The transport of CH3Hg+ across the canalicular membrane has been studied more extensively and is better defined. Findings from numerous studies indicate that transport of CH3Hg+ from hepatocytes into the biliary canaliculus occurs in association with GSH (Refsvik & Norseth, 1975; Ballatori & Clarkson, 1982, 1983, 1985a, Refsvik, 1982). This is not surprising since the majority of CH3Hg+ within hepatocytes appears to be bound to GSH (Omata et al., 1978). Interestingly, increased hepatocellular levels of GSH correspond to a rise in the excretion of GSH and CH3Hg+ into bile (Magos et al., 1978). In contrast, a reduction in the hepatic and biliary levels of GSH corresponds to a reduced accumulation of CH3Hg+ in liver (Refsvik, 1978). Therefore, it seems that the intracellular concentration of GSH significantly impacts the hepatic transport of CH3Hg+. It may be postulated that CH3Hg-S-G, formed within hepatocytes, is transported across the canalicular membrane into bile. Since CH3Hg-S-G is similar structurally to GSH, it is possible that this conjugate may utilize a GSH transporter in the canalicular membrane for export out of hepatocytes (Table 2). Indeed, Dutczak and Ballatori (1994) suggested that a GSH transport system in the canalicular membrane plays a significant role in the biliary secretion of CH3Hg-S-G. MRP2, which is capable of transporting GSH, has since been identified in the canalicular membrane of hepatocytes (Fernandez-Checa et al., 1992, 1993; Garcia-Ruiz et al., 1992; Ballatori & Dutczak, 1994; Ballatori & Truong, 1995) and thus likely plays an important role in the export of CH3Hg+. Indeed, recent studies in TR rats indicate that MRP2 plays a role in the hepatobiliary elimination of mercuric ions following exposure to CH3Hg+ (Table 2) (Madejczyk et al., 2007; Zalups & Bridges, 2009).

After secretion into bile, CH3Hg-S-G appears to be hydrolytically catabolized in a sequential manner by the plasma membrane enzymes γ-glutamyltransferase and cysteinylglycinase to yield CH3Hg-S-Cys, which can be reabsorbed, both by cells lining the bile ducts and by enterocytes in the intestine (Dutczak et al., 1991; Dutczak & Ballatori, 1992; Ballatori, 1994). Though the actual mechanisms involved in the uptake of mercuric ions along the biliary tree have not been determined, it is reasonable to hypothesize that CH3Hg-S-Cys utilizes one or more amino acid transporters in order to gain access to cells. A number of various amino acid transporters, including system L (LAT3) (Babu et al., 2003), are present in the liver and biliary tree (Bode, 2001; Wagner et al., 2001); however, the exact membrane localization of each carrier has not been reported.

Transport of CH3Hg+ in the Placenta

The deleterious effects of CH3Hg+ on fetal development have been recognized widely as one of the most serious toxicological consequences of CH3Hg+ exposure (Matsumoto et al., 1965; Amin-Zaki et al., 1974; Harada, 1978; 1995; Kajiwara & Inouye, 1986; Inouye & Kajiwara, 1988; Kajiwara & Inouye, 1992; Davidson et al., 2008). Following maternal exposure to CH3Hg+, mercuric ions are taken up readily by the placenta and accumulate subsequently in both placental and fetal tissues (Inouye et al., 1985; Inouye & Kajiwara, 1988; Ask et al., 2002). Despite the clinical significance of this area, little is known about the mechanism(s) by which mercuric ions are taken up and transported across the placenta. In vivo studies (Kajiwara et al., 1996) showed that CH3Hg+ is transported across the rat placenta by a neutral amino acid carrier in a time- and dose-dependent manner. Interestingly, co-administration of CH3Hg+ with methionine increased the placental burden of CH3Hg+. It was proposed that this increase in uptake may be the result of the intracellular conversion of methionine to Cys, which may subsequently combine with CH3Hg+ to form a transportable species of CH3Hg+, i.e., CH3Hg-S-Cys. This conjugate may utilize a neutral amino acid carrier such as system L in order to gain access to placenta trophoblasts. Since the two isoforms of system L, LAT1 and LAT2, appear to mediate the transport of CH3Hg-S-Cys in astrocytes and across epithelial cells of the blood–brain barrier (Aschner et al., 1990; Kerper et al., 1992; Mokrzan et al., 1995; Simmons-Willis et al., 2002), it is logical to postulate that this same carrier may also be involved in the transport of CH3Hg-S-Cys across the placenta (Table 2). In the placenta, LAT1 is localized in the apical (maternal) plasma membrane of trophoblasts while LAT2 is found in the basolateral (fetal) membrane (Kudo & Boyd, 2002). A number of other carrier systems (e.g., amino acid, organic anion) are present in the placenta (Leazer & Klaassen, 2003), and although the roles of these other transporters in the transport of CH3Hg+ have not been examined, they should be considered as possible mechanisms for this transport (Table 2).

The mechanisms on the basolateral membrane of trophoblasts that mediate the uptake of mercuric ions from fetal circulation into the placenta have not been identified. One possible mechanism is OAT4, which is localized in the basolateral membrane of placental trophoblasts (Table 2) (Cha et al., 2000; St-Pierre et al., 2000). Since other members of the OAT family transport mercuric ions, it is possible that OAT4 may also be capable of mediating the transport of mercuric species. To date, the ability of OAT4 to transport mercuric ions has not been reported.

Recently, the ability of different chelators to extract mercuric ions from placental and fetal tissues was examined. In pregnant rats exposed to CH3Hg+, treatment with NAC, DMPS, or DMSA facilitates the extraction of mercuric ions from fetal and placental tissues (Aremu et al., 2008; Bridges et al., 2009). OAT4 may be responsible for mediating the transport of mercuric species from fetal tissues, across the basolateral membrane of placental trophoblasts into the intracellular compartments of these cells. On the apical membrane, MRP2, which is localized in the apical membrane of trophoblasts (St-Pierre et al., 2000), may mediate the placental to maternal transfer of mercuric ions (Table 2). Currently, there are no direct data to support these theories.

CONCLUSIONS

In writing this review, every effort has been made to summarize the current literature related to the transport of mercuric ions in various organ systems and tissues. Published studies provide strong evidence for the involvement of amino acid, anion, and drug transporters in the uptake and secretion of mercuric ions in various organs and tissues. Despite the growing body of literature pertaining to the handling of various species of mercury within the body, there remain numerous gaps in our knowledge. This lack of knowledge may be due, in part, to the difficulties associated with working with mercuric species. For example, the chemical bonding nature of mercury often complicates experiments because of bonding to plasma membranes and intracellular proteins.

Not surprisingly, the majority of research to date focused on mechanisms by which Hg2+ and CH3Hg+ are transported within their target organs, i.e., the kidney and brain, respectively. In contrast, little is known about the mechanisms by which mercuric ions are transported by cells of other organs. Although it is clear that intestinal absorption of Hg2+ and CH3Hg+ occurs, the exact mechanisms involved in the uptake and secretion of mercuric ions by enterocytes have not been identified. In addition, hepatic handling of mercuric species has not been elucidated fully. Mechanisms for the handling of mercuric ions have been identified on the canalicular membrane of hepatocytes, yet the mechanisms that allow mercuric ions to enter hepatocytes at the sinusoidal membrane remain unclear. Furthermore, little is known about the transport of mercuric ions across the placenta, despite the clinical significance of this area of research.

Given the lack of clarity regarding the movement of mercuric ions across cellular plasma membranes, it is clear that a great deal of research is still required in order to understand completely the mechanisms involved in the transport of mercury. Experiments utilizing models such as knockout animals, transfected cells, membrane vesicles, and Xenopus laevis oocytes expressing specific transporters will be necessary to characterize the mechanisms involved in the uptake and secretion of mercuric species by cells of various organs. A thorough understanding of both these mechanisms and the way in which mercury is handled at the cellular and molecular levels may lead to advances in treatment regimes for patients intoxicated by mercury, as well a better understanding of the nature and function of various transport proteins of vital substrates.

Acknowledgments

This work was supported by grants from the National Institutes of Health (National Institute of Environmental Health Sciences) awarded to C. C. Bridges (ES015511) and R. K. Zalups (ES05980 and ES11288).

REFERENCES

  1. Adam KR 1951. The effects of dithiols on the distribution of mercury in rabbits. Br. J. Pharmacol. Chemother 6:483–491. [DOI] [PMC free article] [PubMed] [Google Scholar]
  2. Afonso JF and De Alvarez RR 1960. Effects of mercury on human gestation. Am. J. Obstet. Gynecol 80:145–154. [DOI] [PubMed] [Google Scholar]
  3. Akerboom TP, Narayanaswami V, Kunst M, and Sies H 1991. ATP-dependent S-2,4-dinitrophenylglutathione transport in canalicular plasma membrane vesicles from rat liver. J. Biol. Chem 266:13147–13152. [PubMed] [Google Scholar]
  4. Alexander J, and Aaseth J 1982. Organ distribution and cellular uptake of methyl mercury in the rat as influenced by the intra-and extracellular glutathione concentration. Biochem. Pharmacol 31:685–690. [DOI] [PubMed] [Google Scholar]
  5. Amin-Zaki L, Elhassani S, Majeed MA, Clarkson TW, Doherty RA, and Greenwood M 1974. Intra-uterine methylmercury poisoning in Iraq. Pediatrics 54:587–595. [PubMed] [Google Scholar]
  6. Andres S, Laporte JM, and Mason RP 2002. Mercury accumulation and flux across the gills and the intestine of the blue crab Callinectes sapidus. Aquat. Toxicol 56:303–320. [DOI] [PubMed] [Google Scholar]
  7. Aposhian HV 1983. DMSA and DMPS—Water soluble antidotes for heavy metal poisoning. Annu. Rev. Pharmacol. Toxicol 23:193–215. [DOI] [PubMed] [Google Scholar]
  8. Aposhian HV, Maiorino RM, Rivera M, Bruce DC, Dart RC, Hurlbut KM, Levine DJ, Zheng W, Fernando Q, Carter D, and Aposhian MM 1992. Human studies with the chelating agents, DMPS and DMSA. J. Toxicol. Clin. Toxicol 30:505–528. [DOI] [PubMed] [Google Scholar]
  9. Aremu DA, Madejczyk MS, and Ballatori N 2008. N-Acetylcysteine as a potential antidote and biomonitoring agent of methylmercury exposure. Environ. Health Perspect 116:26–31. [DOI] [PMC free article] [PubMed] [Google Scholar]
  10. Aschner M, and Clarkson TW 1988. Uptake of methylmercury in the rat brain: Effects of amino acids. Brain Res. 462:31–39. [DOI] [PubMed] [Google Scholar]
  11. Aschner M, and Clarkson TW 1989. Methyl mercury uptake across bovine brain capillary endothelial cells in vitro: the role of amino acids. Pharmacol. Toxicol 64:293–297. [DOI] [PubMed] [Google Scholar]
  12. Aschner M, Eberle NB, Goderie S, and Kimelberg HK 1990. Methylmercury uptake in rat primary astrocyte cultures: The role of the neutral amino acid transport system. Brain Res. 521:221–228. [DOI] [PubMed] [Google Scholar]
  13. Aschner M, Eberle NB, and Kimelberg HK 1991. Interactions of methylmercury with rat primary astrocyte cultures: Methylmercury efflux. Brain Res. 554:10–14. [DOI] [PubMed] [Google Scholar]
  14. Ashe WF, Largent EJ, Dutra FR, Hubbard DM, and Blackstone M 1953. Behavior of mercury in the animal organism following inhalation. AMA Arch. Ind. Hyg. Occup. Med 7:19–43. [PubMed] [Google Scholar]
  15. Ask K, Akesson A, Berglund M, and Vahter M 2002. Inorganic mercury and methylmercury in placentas of Swedish women. Environ. Health Perspect 110:523–526. [DOI] [PMC free article] [PubMed] [Google Scholar]
  16. Aslamkhan AG, Han YH, Yang XP, Zalups RK, and Pritchard JB 2003. Human renal organic anion transporter 1-dependent uptake and toxicity of mercuricthiol conjugates in Madin-Darby canine kidney cells. Mol. Pharmacol 63:590–596. [DOI] [PubMed] [Google Scholar]
  17. ATSDR. 2007. Toxicological profile for mercury. Atlanta, GA: U.S. Department of Health and Human Services, Centers for Disease Control. [Google Scholar]
  18. Babu E, Kanai Y, Chairoungdua A, Kim DK, Iribe Y, Tangtrongsup S, Jutabha P, Li Y, Ahmed N, Sakamoto S, Anzai N, Nagamori S, and Endou H 2003. Identification of a novel system L amino acid transporter structurally distinct from heterodimeric amino acid transporters. J. Biol. Chem 278:43838–43845. [DOI] [PubMed] [Google Scholar]
  19. Baggett JM, and Berndt WO 1986. The effect of depletion of nonprotein sulfhydryls by diethyl maleate plus buthionine sulfoximine on renal uptake of mercury in the rat. Toxicol. Appl. Pharmacol 83:556–562. [DOI] [PubMed] [Google Scholar]
  20. Bahn A, Hagos Y, Reuter S, Balen D, Brzica H, Krick W, Burckhardt BC, Sabolic I, and Burckhardt G 2008. Identification of a new urate and high affinity nicotinate transporter, hOAT10 SLC22A13. J. Biol. Chem 283:16332–16341. [DOI] [PubMed] [Google Scholar]
  21. Bahn A, Knabe M, Hagos Y, Rodiger M, Godehardt S, Graber-Neufeld DS, Evans KK, Burckhardt G, and Wright SH 2002. Interaction of the metal chelator 2,3-dimercapto-1-propanesulfonate with the rabbit multispecific organic anion transporter 1 rbOAT1. Mol. Pharmacol 62:1128–1136. [DOI] [PubMed] [Google Scholar]
  22. Ballatori N 1994. Glutathione mercaptides as transport forms of metals. Adv. Pharmacol 27:271–298. [DOI] [PubMed] [Google Scholar]
  23. Ballatori N, and Clarkson TW 1982. Developmental changes in the biliary excretion of methylmercury and glutathione. Science 216:61–63. [DOI] [PubMed] [Google Scholar]
  24. Ballatori N, and Clarkson TW 1983. Biliary transport of glutathione and methyl mercury. Am. J. Physiol 244:G435–G441. [DOI] [PubMed] [Google Scholar]
  25. Ballatori N, and Clarkson TW 1984. Dependence of biliary secretion of inorganic mercury on the biliary transport of glutathione. Biochem. Pharmacol 33:1093–1098. [DOI] [PubMed] [Google Scholar]
  26. Ballatori N, and Clarkson TW 1985a. Biliary secretion of glutathione and of glutathione-metal complexes. Fundam. Appl. Toxicol 5:816–831. [DOI] [PubMed] [Google Scholar]
  27. Ballatori N, and Clarkson TW 1985b. Sulfobromophthalein inhibition of glutathione and methylmercury secretion into bile. Am. J. Physiol 248:G238–G245. [DOI] [PubMed] [Google Scholar]
  28. Ballatori N, and Dutczak WJ 1994. Identification and characterization of high and low affinity transport systems for reduced glutathione in liver cell canalicular membranes. J. Biol. Chem 269:19731–19737. [PubMed] [Google Scholar]
  29. Ballatori N, Hammond CL, Cunningham JB, Krance SM, and Marchan R 2005. Molecular mechanisms of reduced glutathione transport: Role of the MRP/CFTR/ABCC and OATP/SLC21A families of membrane proteins. Toxicol. Appl. Pharmacol 204:238–255. [DOI] [PubMed] [Google Scholar]
  30. Ballatori N, and Truong AT 1995. Multiple canalicular transport mechanisms for glutathione S-conjugates. Transport on both ATP- and voltage-dependent carriers. J. Biol. Chem 270:3594–3601. [DOI] [PubMed] [Google Scholar]
  31. Ballatori N, Wang W, and Lieberman MW 1998. Accelerated methylmercury elimination in gamma-glutamyl transpeptidase-deficient mice. Am. J. Pathol 152:1049–1055. [PMC free article] [PubMed] [Google Scholar]
  32. Bayen S, Koroleva E, Lee HK, and Obbard JP 2005. Persistent organic pollutants and heavy metals in typical seafoods consumed in Singapore. J. Toxicol. Environ. Health A 68:151–166. [DOI] [PubMed] [Google Scholar]
  33. Berlin M, and Gibson S 1963. Renal uptake, excretion, and retention of mercury. I. A study in the rabbit during infusion of mercuric chloride. Arch. Environ. Health 6:617–625. [DOI] [PubMed] [Google Scholar]
  34. Berlin M, and Ullberg S 1963a. Accumulation and retention of mercury in the mouse. I. An autoradiographic study after a single intravenous injection of mercuric chloride. Arch. Environ. Health 6:589–601. [DOI] [PubMed] [Google Scholar]
  35. Berlin M, and Ullberg S, 1963b. Accumulation and retention of mercury in the mouse. II. An autoradiographic comparison of phenylmercuric acetate with inorganic mercury. Arch. Environ. Health 6:602–609. [DOI] [PubMed] [Google Scholar]
  36. Berlin M, and Ullberg S 1963c. Accumulation and retention of mercury in the mouse. III. An autoradiographic comparison of methylmercuric dicyandiamide with inorganic mercury. Arch. Environ. Health 6:610–616. [DOI] [PubMed] [Google Scholar]
  37. Berndt WO, Baggett JM, Blacker A, and Houser M 1985. Renal glutathione and mercury uptake by kidney. Fundam. Appl. Toxicol 5:832–839. [DOI] [PubMed] [Google Scholar]
  38. Betz AL, and Goldstein GW 1978. Polarity of the blood–brain barrier: Neutral amino acid transport into isolated brain capillaries. Science 202:225–227. [DOI] [PubMed] [Google Scholar]
  39. Bluhm RE, Breyer JA, Bobbitt RG, Welch LW, Wood AJ, and Branch RA 1992. Elemental mercury vapour toxicity, treatment, and prognosis after acute, intensive exposure in chloralkali plant workers. Part II: Hyperchloraemia and genitourinary symptoms. Hum. Exp. Toxicol 11:211–215. [DOI] [PubMed] [Google Scholar]
  40. Bode BP 2001. Recent molecular advances in mammalian glutamine transport. J. Nutr 131:2475S–2485S; discussion 2486S-2477S. [DOI] [PubMed] [Google Scholar]
  41. Bridges CC, Bauch C, Verrey F, and Zalups RK 2004. Mercuric conjugates of cysteine are transported by the amino acid transporter system b0,+: implications of molecular mimicry. J. Am. Soc. Nephrol 15:663–673. [DOI] [PMC free article] [PubMed] [Google Scholar]
  42. Bridges CC, Joshee L, and Zalups RK 2008a. MRP2 and the DMPS- and DMSA-mediated elimination of mercury in TR- and control rats exposed to thiol S-conjugates of inorganic mercury. Toxicol. Sci 105:211–220. [DOI] [PMC free article] [PubMed] [Google Scholar]
  43. Bridges CC, Joshee L, and Zalups RK 2008b. Multidrug resistance proteins and the renal elimination of inorganic mercury mediated by 2,3-dimercaptopropane-1-sulfonic acid and meso-2,3-dimercaptosuccinic acid. J. Pharmacol. Exp. Ther 324:383–390. [DOI] [PMC free article] [PubMed] [Google Scholar]
  44. Bridges CC, Joshee L, and Zalups RK 2009. Effect of DMPS and DMSA on the placental and fetal disposition of methylmercury. Placenta 30:800–805. [DOI] [PMC free article] [PubMed] [Google Scholar]
  45. Bridges CC, and Zalups RK 2004. Homocysteine, system b0,+ and the renal epithelial transport and toxicity of inorganic mercury. Am. J. Pathol 165:1385–1394. [DOI] [PMC free article] [PubMed] [Google Scholar]
  46. Bridges CC, and Zalups RK 2006. System b0,+ and the transport of thiol-S-conjugates of methylmercury. J. Pharmacol. Exp. Ther 319:948–956. [DOI] [PubMed] [Google Scholar]
  47. Buchler M, Konig J, Brom M, Kartenbeck J, Spring H, Horie T, and Keppler D 1996. cDNA cloning of the hepatocyte canalicular isoform of the multidrug resistance protein, cMrp, reveals a novel conjugate export pump deficient in hyperbilirubinemic mutant rats. J. Biol. Chem 271:15091–15098. [DOI] [PubMed] [Google Scholar]
  48. Burckhardt BC, Drinkuth B, Menzel C, Konig A, Steffgen J, Wright SH, and Burckhardt G 2002. The renal Na+-dependent dicarboxylate transporter, NaDC-3, translocates dimethyl- and disulfhydryl-compounds and contributes to renal heavy metal detoxification. J. Am. Soc. Nephrol 13:2628–2638. [DOI] [PubMed] [Google Scholar]
  49. Cannon VT, Barfuss DW, and Zalups RK 2000. Molecular homology and the luminal transport of Hg2+ in the renal proximal tubule. J. Am. Soc. Nephrol 11:394–402. [DOI] [PubMed] [Google Scholar]
  50. Cannon VT, Zalups RK, and Barfuss DW 2001. Amino acid transporters involved in luminal transport of mercuric conjugates of cysteine in rabbit proximal tubule. J. Pharmacol. Exp. Ther 298:780–789. [PubMed] [Google Scholar]
  51. Canonne-Hergaux F, Gruenheid S, Ponka P, and Gros P 1999. Cellular and subcellular localization of the Nramp2 iron transporter in the intestinal brush border and regulation by dietary iron. Blood 93:4406–4417. [PubMed] [Google Scholar]
  52. Carmignani M, Boscolo P, Artese L, Del Rosso G, Porcelli G, Felaco M, Volpe AR, and Giuliano G 1992. Renal mechanisms in the cardiovascular effects of chronic exposure to inorganic mercury in rats. Br. J. Ind. Med 49:226–232. [DOI] [PMC free article] [PubMed] [Google Scholar]
  53. Cember H, Gallagher P, and Faulkner A 1968. Distribution of mercury among blood fractions and serum proteins. Am. Ind. Hyg. Assoc. J 29:233–237. [DOI] [PubMed] [Google Scholar]
  54. Cha SH, Sekine T, Kusuhara H, Yu E, Kim JY, Kim DK, Sugiyama Y, Kanai Y, and Endou H 2000. Molecular cloning and characterization of multispecific organic anion transporter 4 expressed in the placenta. J. Biol. Chem 275:4507–4512. [DOI] [PubMed] [Google Scholar]
  55. Cherian MG, and Clarkson TW 1976. Biochemical changes in rat kidney on exposure to elemental mercury vapor: Effect on biosynthesis of metallothionein. Chem. Biol. Interact 12:109–120. [DOI] [PubMed] [Google Scholar]
  56. Chillaron J, Roca R, Valencia A, Zorzano A, and Palacin M 2001. Heteromeric amino acid transporters: biochemistry, genetics, and physiology. Am. J. Physiol. Renal Physiol 281:F995–F1018. [DOI] [PubMed] [Google Scholar]
  57. Clarkson TW 1993. Molecular and ionic mimicry of toxic metals. Annu. Rev. Pharmacol. Toxicol 33:545–571. [DOI] [PubMed] [Google Scholar]
  58. Clarkson TW, and Magos L 1966. Studies on the binding of mercury in tissue homogenates. Biochem. J 99:62–70. [DOI] [PMC free article] [PubMed] [Google Scholar]
  59. Clarkson TW, and Magos L 2006. The toxicology of mercury and its chemical compounds. Crit. Rev. Toxicol 36:609–662. [DOI] [PubMed] [Google Scholar]
  60. Dalton TP, He L, Wang B, Miller ML, Jin L, Stringer KF, Chang X, Baxter CS, and Nebert DW 2005. Identification of mouse SLC39A8 as the transporter responsible for cadmium-induced toxicity in the testis. Proc. Natl. Acad. Sci. USA 102:3401–3406. [DOI] [PMC free article] [PubMed] [Google Scholar]
  61. Danscher G, Horsted-Bindslev P, and Rungby J 1990. Traces of mercury in organs from primates with amalgam fillings. Exp. Mol. Pathol 52:291–299. [DOI] [PubMed] [Google Scholar]
  62. Dave MH, Schulz N, Zecevic M, Wagner CA, and Verrey F 2004. Expression of heteromeric amino acid transporters along the murine intestine. J. Physiol 558:597–610. [DOI] [PMC free article] [PubMed] [Google Scholar]
  63. Davidson PW, Strain JJ, Myers GJ, Thurston SW, Bonham MP, Shamlaye CF, Stokes-Riner A, Wallace JM, Robson PJ, Duffy EM, Georger LA, Sloane-Reeves J, Cernichiari E, Canfield RL, Cox C, Huang LS, Janciuras J, and Clarkson TW 2008. Neurodevelopmental effects of maternal nutritional status and exposure to methylmercury from eating fish during pregnancy. Neurotoxicology 29:767–775. [DOI] [PMC free article] [PubMed] [Google Scholar]
  64. de Ceaurriz J, and Ban M 1990. Role of gamma-glutamyltranspeptidase and betalyase in the nephrotoxicity of hexachloro-1,3-butadiene and methyl mercury in mice. Toxicol. Lett 50:249–256. [DOI] [PubMed] [Google Scholar]
  65. de Ceaurriz J, Payan JP, Morel G, and Brondeau MT 1994. Role of extracellular glutathione and gamma-glutamyltranspeptidase in the disposition and kidney toxicity of inorganic mercury in rats. J. Appl. Toxicol 14:201–206. [DOI] [PubMed] [Google Scholar]
  66. Deeley RG, Westlake C, and Cole SP 2006. Transmembrane transport of endo- and xenobiotics by mammalian ATP-binding cassette multidrug resistance proteins. Physiol. Rev 86:849–899. [DOI] [PubMed] [Google Scholar]
  67. Di Simplicio P, Gorelli M, Ciuffreda P, and Leonzio C 1990. The relationship between gamma-glutamyl transpeptidase and Hg levels in Se/Hg antagonism in mouse liver and kidney. Pharmacol. Res 22:515–526. [DOI] [PubMed] [Google Scholar]
  68. Diamond GL, Klotzbach JM, and Stewart JR 1988. Complexing activity of 2,3-dimercapto-1-propanesulfonate and its disulfide auto-oxidation product in rat kidney. J. Pharmacol. Exp. Ther 246:270–274. [PubMed] [Google Scholar]
  69. Dietert RR, and Dietert JM 2008. Potential for early-life immune insult including developmental immunotoxicity in autism and autism spectrum disorders: Focus on critical windows of immune vulnerability. J. Toxicol. Environ. Health B 11:660–680. [DOI] [PubMed] [Google Scholar]
  70. Dunn JD, and Clarkson TW 1980. Does mercury exhalation signal demethylation of methylmercury? Health Phys. 38:411–414. [PubMed] [Google Scholar]
  71. Dutczak WJ, and Ballatori N 1992. Gamma-glutamyltransferase-dependent biliary-hepatic recycling of methyl mercury in the guinea pig. J. Pharmacol. Exp. Ther 262:619–623. [PubMed] [Google Scholar]
  72. Dutczak WJ, and Ballatori N 1994. Transport of the glutathione–methylmercury complex across liver canalicular membranes on reduced glutathione carriers. J. Biol. Chem 269:9746–9751. [PubMed] [Google Scholar]
  73. Dutczak WJ, Clarkson TW, and Ballatori N 1991. Biliary-hepatic recycling of a xenobiotic: gallbladder absorption of methyl mercury. Am. J. Physiol 260:G873–G880. [DOI] [PubMed] [Google Scholar]
  74. Endo T, Nakaya S, Kimura R, and Murata T 1984. Gastrointestinal absorption of inorganic mercuric compounds in vivo and in situ. Toxicol. Appl. Pharmacol 74:223–229. [DOI] [PubMed] [Google Scholar]
  75. Fernandez-Checa JC, Ookhtens M, and Kaplowitz N 1993. Selective induction by phenobarbital of the electrogenic transport of glutathione and organic anions in rat liver canalicular membrane vesicles. J. Biol. Chem 268:10836–10841. [PubMed] [Google Scholar]
  76. Fernandez-Checa JC, Takikawa H, Horie T, Ookhtens M, and Kaplowitz N 1992. Canalicular transport of reduced glutathione in normal and mutant Eisai hyperbilirubinemic rats. J. Biol. Chem 267:1667–1673. [PubMed] [Google Scholar]
  77. Ferrier B, Martin M, and Roch-Ramel F 1983. Effects of p-aminohippurate and pyrazinoate on the renal excretion of salicylate in the rat: A micropuncture study. J. Pharmacol. Exp. Ther 224:451–458. [PubMed] [Google Scholar]
  78. Foulkes EC 1993. Metallothionein and glutathione as determinants of cellular retention and extrusion of cadmium and mercury. Life Sci. 52:1617–1620. [DOI] [PubMed] [Google Scholar]
  79. Foulkes EC 2000. Transport of toxic heavy metals across cell membranes. Proc. Soc. Exp. Biol. Med 223:234–240. [DOI] [PubMed] [Google Scholar]
  80. Foulkes EC, and Bergman D 1993. Inorganic mercury absorption in mature and immature rat jejunum: Transcellular and intercellular pathways in vivo and in everted sacs. Toxicol. Appl. Pharmacol 120:89–95. [DOI] [PubMed] [Google Scholar]
  81. Friberg L 1959. Studies on the metabolism of mercuric chloride and methyl mercury dicyandiamide; Experiments on rats given subcutaneous injections with radioactive mercury Hg203. AMA Arch. Ind. Health 20:42–49. [PubMed] [Google Scholar]
  82. Friberg L, Odeblad E, and Forssman S 1957. Distribution of two mercury compounds in rabbits after a single subcutaneous injection; a radiometric and autoradiographic study of the distribution of mercuric chloride and phenylmercuric acetate. AMA Arch. Ind. Health 16:163–168. [PubMed] [Google Scholar]
  83. Friedman HL 1957. Relationship between chemical structure and biological activity in mercurial compounds. Ann. N Y Acad. Sci 65:461–470. [DOI] [PubMed] [Google Scholar]
  84. Fuhr BJ, and Rabenstein DL 1973. Nuclear magnetic resonance studies of the solution chemistry of metal complexes. IX. The binding of cadmium, zinc, lead, and mercury by glutathione. J. Am. Chem. Soc 95:6944–6950. [DOI] [PubMed] [Google Scholar]
  85. Gage JC 1964. Distribution and excretion of methyl and phenyl mercury salts. Br. J. Ind. Med 21:197–202. [DOI] [PMC free article] [PubMed] [Google Scholar]
  86. Ganapathy V, Ganapathy ME, and Leibach FH 2001. Intestinal transport of peptides and amino acids. New York: Academic Press. [Google Scholar]
  87. Garcia-Ruiz C, Fernandez-Checa JC, and Kaplowitz N 1992. Bidirectional mechanism of plasma membrane transport of reduced glutathione in intact rat hepatocytes and membrane vesicles. J. Biol. Chem 267:22256–22264. [PubMed] [Google Scholar]
  88. Geier DA, and Geier MR 2007. A prospective study of mercury toxicity biomarkers in autistic spectrum disorders. J. Toxicol. Environ. Health A 70:1723–1730. [DOI] [PubMed] [Google Scholar]
  89. Gonska T, Hirsch JR, and Schlatter E 2000. Amino acid transport in the renal proximal tubule. Amino Acids 19:395–407. [DOI] [PubMed] [Google Scholar]
  90. Gregus Z, Stein AF, and Klaassen CD 1987. Effect of inhibition of gamma-glutamyltranspeptidase on biliary and urinary excretion of glutathione-derived thiols and methylmercury. J. Pharmacol. Exp. Ther 242:27–32. [PubMed] [Google Scholar]
  91. Hahn LJ, Kloiber R, Leininger RW, Vimy MJ, and Lorscheider FL 1990. Whole-body imaging of the distribution of mercury released from dental fillings into monkey tissues. FASEB J. 4:3256–3260. [DOI] [PubMed] [Google Scholar]
  92. Hahn LJ, Kloiber R, Vimy MJ, Takahashi Y, and Lorscheider FL 1989. Dental “silver” tooth fillings: A source of mercury exposure revealed by whole-body image scan and tissue analysis. FASEB J. 3:2641–2646. [DOI] [PubMed] [Google Scholar]
  93. Harada M 1978. Congenital Minamata disease: Intrauterine methylmercury poisoning. Teratology 18:285–288. [DOI] [PubMed] [Google Scholar]
  94. Harada M 1995. Minamata disease: Methylmercury poisoning in Japan caused by environmental pollution. Crit. Rev. Toxicol 25:1–24. [DOI] [PubMed] [Google Scholar]
  95. He L, Girijashanker K, Dalton TP, Reed J, Li H, Soleimani M, and Nebert DW 2006. ZIP8, member of the solute-carrier-39 SLC39 metal-transporter family: Characterization of transporter properties. Mol. Pharmacol 70:171–180. [DOI] [PubMed] [Google Scholar]
  96. Hilgendorf C, Ahlin G, Seithel A, Artursson P, Ungell AL, and Karlsson J 2007. Expression of thirty-six drug transporter genes in human intestine, liver, kidney, and organotypic cell lines. Drug Metab. Dispos 35:1333–1340. [DOI] [PubMed] [Google Scholar]
  97. Hirayama K 1975. Transport mechanism of methyl mercury. Intestinal absorption, biliary excretion and distribution of methyl mercury. Kumamoto Med. J 28:151–163. [PubMed] [Google Scholar]
  98. Hirayama K 1980. Effect of amino acids on brain uptake of methyl mercury. Toxicol. Appl. Pharmacol 55:318–323. [DOI] [PubMed] [Google Scholar]
  99. Hirayama K 1985. Effects of combined administration of thiol compounds and methylmercury chloride on mercury distribution in rats. Biochem. Pharmacol 34:2030–2032. [DOI] [PubMed] [Google Scholar]
  100. Hoyle I, and Handy RD 2005. Dose-dependent inorganic mercury absorption by isolated perfused intestine of rainbow trout, Oncorhynchus mykiss, involves both amiloride-sensitive and energy-dependent pathways. Aquat. Toxicol 72:147–159. [DOI] [PubMed] [Google Scholar]
  101. Hughes WL 1957. A physicochemical rationale for the biological activity of mercury and its compounds. Ann. NY Acad. Sci 65:454–460. [DOI] [PubMed] [Google Scholar]
  102. Hultman P, Bell LJ, Enestrom S, and Pollard KM 1992. Murine susceptibility to mercury. I. Autoantibody profiles and systemic immune deposits in inbred, congenic, and intra-H-2 recombinant strains. Clin. Immunol. Immunopathol 65:98–109. [DOI] [PubMed] [Google Scholar]
  103. Hultman P, and Enestrom S 1986. Localization of mercury in the kidney during experimental acute tubular necrosis studied by the cytochemical silver amplification method. Br. J. Exp. Pathol 67:493–503. [PMC free article] [PubMed] [Google Scholar]
  104. Hultman P, and Enestrom S 1992. Dose-response studies in murine mercury-induced autoimmunity and immune-complex disease. Toxicol. Appl. Pharmacol 113:199–208. [DOI] [PubMed] [Google Scholar]
  105. Hultman P, Enestrom S, and von Schenck H 1985. Renal handling of inorganic mercury in mice. The early excretion phase following a single intravenous injection of mercuric chloride studied by the silver amplification method. Virchows Arch. B Cell. Pathol. Incl. Mol. Pathol 49:209–224. [PubMed] [Google Scholar]
  106. Iioka H, Moriyama I, Oku M, Hino K, Itani Y, Okamura Y, and Ichijo M 1987. [The effect of inorganic mercury on placental amino acid transport using microvillous membrane vesicles]. Nippon Sanka Fujinka Gakkai Zasshi 39:202–206. [PubMed] [Google Scholar]
  107. Ilback NG, Frisk P, Tallkvist J, Gadhasson IL, Blomberg J, and Friman G 2008. Gastrointestinal uptake of trace elements are changed during the course of a common human viral Coxsackievirus B3 infection in mice. J. Trace Elem. Med. Biol 22:120–130. [DOI] [PubMed] [Google Scholar]
  108. Inouye M, and Kajiwara Y 1988. Developmental disturbances of the fetal brain in guinea-pigs caused by methylmercury. Arch. Toxicol 62:15–21. [DOI] [PubMed] [Google Scholar]
  109. Inouye M, and Kajiwara Y 1990. Strain difference of the mouse in manifestation of hydrocephalus following prenatal methylmercury exposure. Teratology 41:205–210. [DOI] [PubMed] [Google Scholar]
  110. Inouye M, Murao K, and Kajiwara Y 1985. Behavioral and neuropathological effects of prenatal methylmercury exposure in mice. Neurobehav. Toxicol. Teratol 7:227–232. [PubMed] [Google Scholar]
  111. Islinger F, Gekle M, and Wright SH 2001. Interaction of 2,3-dimercapto-1-propane sulfonate with the human organic anion transporter hOAT1. J. Pharmacol. Exp. Ther 299:741–747. [PubMed] [Google Scholar]
  112. Itagaki S, Chiba M, Kobayashi M, Hirano T, and Iseki K 2008. Contribution of multidrug resistance-associated protein 2 to secretory intestinal transport of organic anions. Biol. Pharm. Bull 31:146–148. [DOI] [PubMed] [Google Scholar]
  113. Jaffe KM, Shurtleff DB, and Robertson WO 1983. Survival after acute mercury vapor poisoning. Am. J. Dis. Child 137:749–751. [DOI] [PubMed] [Google Scholar]
  114. Jansson T 2001. Amino acid transporters in the human placenta. Pediatr. Res 49:141–147. [DOI] [PubMed] [Google Scholar]
  115. Jernelov AA 1973. A new biochemical pathway for the methylation of mercury and some ecolgical implications. Springfield, IL: Thomas. [Google Scholar]
  116. Joshi JG, Sczekan SR, and Fleming JT 1989. Ferritin—A general metal detoxicant. Biol. Trace Elem. Res 21:105–110. [DOI] [PubMed] [Google Scholar]
  117. Jung SM, Lee WK, Kwak JO, Jung SY, Park J, Kim WY, Kim J, and Cha SH 2006. Identification of a novel murine organic anion transporter like protein 1 OATLP1 expressed in the kidney. Exp. Mol. Med 38:485–493. [DOI] [PubMed] [Google Scholar]
  118. Kajiwara Y, and Inouye M 1986. Effects of methylmercury and mercuric chloride on preimplantation mouse embryos in vivo. Teratology 33:231–237. [DOI] [PubMed] [Google Scholar]
  119. Kajiwara Y, and Inouye M 1992. Inhibition of implantation caused by methylmercury and mercuric chloride in mouse embryos in vivo. Bull. Environ. Contam. Toxicol 49:541–546. [DOI] [PubMed] [Google Scholar]
  120. Kajiwara Y, Yasutake A, Adachi T, and Hirayama K 1996. Methylmercury transport across the placenta via neutral amino acid carrier. Arch. Toxicol 70:310–314. [DOI] [PubMed] [Google Scholar]
  121. Keppler D, and Konig J 1997. Hepatic canalicular membrane 5: Expression and localization of the conjugate export pump encoded by the MRP2 cMRP/cMOAT gene in liver. FASEB J. 11:509–516. [DOI] [PubMed] [Google Scholar]
  122. Kern JK, and Jones AM 2006. Evidence of toxicity, oxidative stress, and neuronal insult in autism. J. Toxicol. Environ. Health B 9:485–499. [DOI] [PubMed] [Google Scholar]
  123. Kerper LE, Ballatori N, and Clarkson TW 1992. Methylmercury transport across the blood-brain barrier by an amino acid carrier. Am. J. Physiol 262:R761–R765. [DOI] [PubMed] [Google Scholar]
  124. Kitamura Y, Kusuhara H, and Sugiyama Y 2009. Functional characterization of multidrug resistance-associated protein 3 Mrp3/Abcc3 in the basolateral efflux of glucuronide conjugates in the mouse small intestine. J. Pharmacol. Exp. Ther Epub ahead of print: PMID: 19889793. [DOI] [PubMed] [Google Scholar]
  125. Koh AS, Simmons-Willis TA, Pritchard JB, Grassl SM, and Ballatori N 2002. Identification of a mechanism by which the methylmercury antidotes N-acetylcysteine and dimercaptopropanesulfonate enhance urinary metal excretion: Transport by the renal organic anion transporter-1. Mol. Pharmacol 62:921–926. [DOI] [PubMed] [Google Scholar]
  126. Kojima R, Sekine T, Kawachi M, Cha SH, Suzuki Y, and Endou H 2002. Immunolocalization of multispecific organic anion transporters, OAT1, OAT2, and OAT3, in rat kidney. J. Am. Soc. Nephrol 13:848–857. [DOI] [PubMed] [Google Scholar]
  127. Kudo Y, and Boyd CA 2002. Human placental amino acid transporter genes: Expression and function. Reproduction 124:593–600. [DOI] [PubMed] [Google Scholar]
  128. Landner L 1971. Biochemical model for the biological methylation of mercury suggested from methylation studies in vivo with Neurospora crassa. Nature 230:452–454. [DOI] [PubMed] [Google Scholar]
  129. Laporte JM, Andres S, and Mason RP 2002. Effect of ligands and other metals on the uptake of mercury and methylmercury across the gills and the intestine of the blue crab Callinectes sapidus. Comp. Biochem. Physiol. C Toxicol Pharmacol 131:185–196. [DOI] [PubMed] [Google Scholar]
  130. Lau S, and Sarkar B 1979. Inorganic mercury(II)-binding components in normal human blood serum. J. Toxicol. Environ. Health 5:907–916. [DOI] [PubMed] [Google Scholar]
  131. Leaner JJ, and Mason RP 2002. Methylmercury accumulation and fluxes across the intestine of channel catfish, Ictalurus punctatus. Comp. Biochem. Physiol. C Toxicol. Pharmacol 132:247–259. [DOI] [PubMed] [Google Scholar]
  132. Leazer TM, and Klaassen CD 2003. The presence of xenobiotic transporters in rat placenta. Drug Metab. Dispos 31:153–167. [DOI] [PubMed] [Google Scholar]
  133. Leslie EM, Deeley RG, and Cole SP 2005. Multidrug resistance proteins: Role of P-glycoprotein, MRP1, MRP2, and BCRP ABCG2 in tissue defense. Toxicol. Appl. Pharmacol 204:216–237. [DOI] [PubMed] [Google Scholar]
  134. Lin JL, and Lim PS 1993. Massive oral ingestion of elemental mercury. J. Toxicol. Clin. Toxicol 31:487–492. [DOI] [PubMed] [Google Scholar]
  135. Liu Z, Li H, Soleimani M, Girijashanker K, Reed JM, He L, Dalton TP, and Nebert DW 2008. Cd2+ versus Zn2+ uptake by the ZIP8 HCO3-dependent symporter: Kinetics, electrogenicity and trafficking. Biochem. Biophys. Res. Commun 365:814–820. J. Ind. Hyg. Toxicol. 31:190–200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  136. Madejczyk MS, Aremu DA, Simmons-Willis TA, Clarkson TW, and Ballatori N 2007. Accelerated urinary excretion of methylmercury following administration of its antidote N-acetylcysteine requires Mrp2/Abcc2, the apical multidrug resistance-associated protein. J. Pharmacol. Exp. Ther 322:378–384. [DOI] [PubMed] [Google Scholar]
  137. Magos L, and Butler WH 1976. The kinetics of methylmercury administered repeatedly to rats. Arch. Toxicol 35:25–39. [DOI] [PubMed] [Google Scholar]
  138. Magos L, Brown AW, Sparrow S, Bailey E, Snowden RT, and Skipp WR 1985. The comparative toxicology of ethyl- and methylmercury. Arch. Toxicol 57:260–267. [DOI] [PubMed] [Google Scholar]
  139. Magos L, Clarkson TW, and Allen J 1978. The interrelationship between non-protein bound thiols and the biliary excretion of methylmercury. Biochem. Pharmacol 27:2203–2208. [DOI] [PubMed] [Google Scholar]
  140. Magos L, Peristianis GC, Clarkson TW, Brown A, Preston S, and Snowden RT 1981. Comparative study of the sensitivity of male and female rats to methylmercury. Arch. Toxicol 48:11–20. [DOI] [PubMed] [Google Scholar]
  141. Maher JM, Slitt AL, Cherrington NJ, Cheng X, and Klaassen CD 2005. Tissue distribution and hepatic and renal ontogeny of the multidrug resistance-associated protein Mrp family in mice. Drug Metab. Dispos 33:947–955. [DOI] [PubMed] [Google Scholar]
  142. Matsumoto H, Koya G, and Takeuchi T 1965. Fetal Minamata disease. A neuropathological study of two cases of intrauterine intoxication by a methyl mercury compound. J. Neuropathol. Exp. Neurol 24:563–574. [PubMed] [Google Scholar]
  143. McNeil SI, Bhatnagar MK, and Turner CJ 1988. Combined toxicity of ethanol and methylmercury in rat. Toxicology 53:345–363. [DOI] [PubMed] [Google Scholar]
  144. Mokrzan EM, Kerper LE, Ballatori N, and Clarkson TW 1995. Methylmercury-thiol uptake into cultured brain capillary endothelial cells on amino acid system L. J. Pharmacol. Exp. Ther 272:1277–1284. [PubMed] [Google Scholar]
  145. Motohashi H, Sakurai Y, Saito H, Masuda S, Urakami Y, Goto M, Fukatsu A, Ogawa O, and Inui K 2002. Gene expression levels and immunolocalization of organic ion transporters in the human kidney. J. Am. Soc. Nephrol 13:866–874. [DOI] [PubMed] [Google Scholar]
  146. Mulder KM, and Kostyniak PJ 1985. Effect of L-alpha S,5S-alpha-amino-3-chloro-4,5-dihydro-5-isoxazoleacetic acid on urinary excretion of methylmercury in the mouse. J. Pharmacol. Exp. Ther 234:156–160. [PubMed] [Google Scholar]
  147. Murphy MJ, Culliford EJ, and Parsons V 1979. A case of poisoning with mercuric chloride. Resuscitation 7:35–44. [DOI] [PubMed] [Google Scholar]
  148. Mussini E 1958. [Distribution in the organism & diuretic activity of p-chlorobenzoic acid mercury salt.]. Boll. Soc. Ital. Biol. Sper 34:1586–1588. [PubMed] [Google Scholar]
  149. Naganuma A, Oda-Urano N, Tanaka T, and Imura N 1988. Possible role of hepatic glutathione in transport of methylmercury into mouse kidney. Biochem. Pharmacol 37:291–296. [DOI] [PubMed] [Google Scholar]
  150. Nakanishi T, Hatanaka T, Huang W, Prasad PD, Leibach FH, Ganapathy ME, and Ganapathy V 2001. Na+- and Cl-coupled active transport of carnitine by the amino acid transporter ATB0,+ from mouse colon expressed in HRPE cells and Xenopus oocytes. J. Physiol 532:297–304. [DOI] [PMC free article] [PubMed] [Google Scholar]
  151. Norseth T, and Clarkson TW 1970a. Biotransformation of methylmercury salts in the rat studied by specific determination of inorganic mercury. Biochem. Pharmacol 19:2775–2783. [DOI] [PubMed] [Google Scholar]
  152. Norseth T, and Clarkson TW 1970b. Studies on the biotransformation of 203Hg-labeled methyl mercury chloride in rats. Arch. Environ. Health 21:717–727. [DOI] [PubMed] [Google Scholar]
  153. Norseth T, and Clarkson TW 1971. Intestinal transport of 203Hg-labeled methyl mercury chloride. Role of biotransformation in rats. Arch. Environ. Health 22:568–577. [DOI] [PubMed] [Google Scholar]
  154. Oka JA, Christensen MD, and Weigel PH 1989. Hyperosmolarity inhibits galactosyl receptor-mediated but not fluid phase endocytosis in isolated rat hepatocytes. J. Biol. Chem 264:12016–12024. [PubMed] [Google Scholar]
  155. Oldendorf WH 1973. Stereospecificity of blood–brain barrier permeability to amino acids. Am. J. Physiol 224:967–969. [DOI] [PubMed] [Google Scholar]
  156. Omata S, Sakimura K, Ishii T, and Sugano H 1978. Chemical nature of a methylmercury complex with a low molecular weight in the liver cytosol of rats exposed to methylmercury chloride. Biochem. Pharmacol 27:1700–1702. [DOI] [PubMed] [Google Scholar]
  157. Omata S, Sato M, Sakimura K, and Sugano H 1980. Time-dependent accumulation of inorganic mercury in subcellular fractions of kidney, liver, and brain of rats exposed to methylmercury. Arch. Toxicol 44:231–241. [DOI] [PubMed] [Google Scholar]
  158. Palacin M, Estevez R, Bertran J, and Zorzano A 1998. Molecular biology of mammalian plasma membrane amino acid transporters. Physiol. Rev 78:969–1054. [DOI] [PubMed] [Google Scholar]
  159. Palacin M, Fernandez E, Chillaron J, and Zorzano A 2001. The amino acid transport system b0,+ and cystinuria. Mol. Membr. Biol 18:21–26. [PubMed] [Google Scholar]
  160. Planas-Bohne F 1981. The effect of 2,3-dimercaptorpropane-1-sulfonate and dimercaptosuccinic acid on the distribution and excretion of mercuric chloride in rats. Toxicology 19:275–278. [DOI] [PubMed] [Google Scholar]
  161. Prickett CS, Laug EP, and Kunze FM 1950. Distribution of mercury in rats following oral and intravenous administration of mercuric acetate and phenylmercuric acetate. Proc. Soc. Exp. Biol. Med 73:585–588. [DOI] [PubMed] [Google Scholar]
  162. Prime-Chapman HM, Fearn RA, Cooper AE, Moore V, and Hirst BH 2004. Differential multidrug resistance-associated protein 1 through 6 isoform expression and function in human intestinal epithelial Caco-2 cells. J. Pharmacol. Exp. Ther 311:476–484. [DOI] [PubMed] [Google Scholar]
  163. Pritchard JB 1988. Coupled transport of -aminohippurate by rat kidney basolateral membrane vesicles. Am. J. Physiol 255:F597–F604. [DOI] [PubMed] [Google Scholar]
  164. Refsvik T 1978. Excretion of methyl mercury in rat bile: The effect of diethylmaleate, cyclohexene oxide and acrylamide. Acta Pharmacol. Toxicol. Copenh 42:135–141. [DOI] [PubMed] [Google Scholar]
  165. Refsvik T 1982. Excretion of methyl mercury in rat bile: The effect of thioctic acid, thionalide, hexadecyl- and octadecylmercaptoacetate. Acta Pharmacol. Toxicol. Copenh 50:196–205. [DOI] [PubMed] [Google Scholar]
  166. Refsvik T, and Norseth T 1975. Methyl mercuric compounds in rat bile. Acta Pharmacol. Toxicol. Copenh 36:67–78. [DOI] [PubMed] [Google Scholar]
  167. Richardson RJ, and Murphy SD 1975. Effect of glutathione depletion on tissue deposition of methylmercury in rats. Toxicol. Appl. Pharmacol 31:505–519. [DOI] [PubMed] [Google Scholar]
  168. Risher JF, and De Rosa CT 2007. Inorganic: The other mercury. J. Enviro.n Health 70:9–16; discussion 40. [PubMed] [Google Scholar]
  169. Rodier PM, Kates B, and Simons R 1988. Mercury localization in mouse kidney over time: Autoradiography versus silver staining. Toxicol. Appl. Pharmacol 92:235–245. [DOI] [PubMed] [Google Scholar]
  170. Rooney JP 2007. The role of thiols, dithiols, nutritional factors and interacting ligands in the toxicology of mercury. Toxicology 234:145–156. [DOI] [PubMed] [Google Scholar]
  171. Rost D, Mahner S, Sugiyama Y, and Stremmel W 2002. Expression and localization of the multidrug resistance-associated protein 3 in rat small and large intestine. Am. J. Physiol. Gastrointest Liver Physiol 282:G720–G726. [DOI] [PubMed] [Google Scholar]
  172. Rowens B, Guerrero-Betancourt D, Gottlieb CA, Boyes RJ, and Eichenhorn MS 1991. Respiratory failure and death following acute inhalation of mercury vapor. A clinical and histologic perspective. Chest 99:185–190. [DOI] [PubMed] [Google Scholar]
  173. Rubino FM, Verduci C, Giampiccolo R, Pulvirenti S, Brambilla G, and Colombi A 2004. Molecular characterization of homo-and heterodimeric mercuryII-bis-thiolates of some biologically relevant thiols by electrospray ionization and triple quadrupole tandem mass spectrometry. J. Am. Soc. Mass Spectrom 15:288–300. [DOI] [PubMed] [Google Scholar]
  174. Ruprecht J 2008. Dimaval: Scientific product monograph. Berlin, Germany: Heyl Pharmaceuticals. [Google Scholar]
  175. Samuels ER, Heick HM, McLaine PN, and Farant JP 1982. A case of accidental inorganic mercury poisoning. J. Anal. Toxicol 6:120–122. [DOI] [PubMed] [Google Scholar]
  176. Schaub TP, Kartenbeck J, Konig J, Spring H, Dorsam J, Staehler G, Storkel S, Thon WF, and Keppler D 1999. Expression of the MRP2 gene-encoded conjugate export pump in human kidney proximal tubules and in renal cell carcinoma. J. Am. Soc. Nephrol 10:1159–1169. [DOI] [PubMed] [Google Scholar]
  177. Schaub TP, Kartenbeck J, Konig J, Vogel O, Witzgall R, Kriz W, and Keppler D 1997. Expression of the conjugate export pump encoded by the mrp2 gene in the apical membrane of kidney proximal tubules. J. Am. Soc. Nephrol 8:1213–1221. [DOI] [PubMed] [Google Scholar]
  178. Shimomura A, Chonko AM, and Grantham JJ 1981. Basis for heterogeneity of para-aminohippurate secretion in rabbit proximal tubules. Am. J. Physiol 240:F430–F436. [DOI] [PubMed] [Google Scholar]
  179. Shoji T, Suzuki H, Kusuhara H, Watanabe Y, Sakamoto S, and Sugiyama Y 2004. ATP-dependent transport of organic anions into isolated basolateral membrane vesicles from rat intestine. Am. J. Physiol. Gastrointest. Liver Physiol 287:G749–G756. [DOI] [PubMed] [Google Scholar]
  180. Simmonds MP, Haraguchi K, Endo T, Cipriano F, Palumbi SR, and Troisi GM 2002. Human health significance of organochlorine and mercury contaminants in japanese whale meat. J. Toxicol. Environ. Health A 65:1211–1235. [DOI] [PubMed] [Google Scholar]
  181. Simmons-Willis TA, Koh AS, Clarkson TW, and Ballatori N 2002. Transport of a neurotoxicant by molecular mimicry: The methyl mercury-L-cysteine complex is a substrate for human L-type large neutral amino acid transporter LAT 1 and LAT2. Biochem. J 367:239–246. [DOI] [PMC free article] [PubMed] [Google Scholar]
  182. Sloan JL, and Mager S 1999. Cloning and functional expression of a human Na+ and Cl-dependent neutral and cationic amino acid transporter B0+. J. Biol. Chem 274:23740–23745. [DOI] [PubMed] [Google Scholar]
  183. Soni JP, Singhania RU, Bansal A, and Rathi G 1992. Acute mercury vapor poisoning. Indian Pediatr. 29:365–368. [PubMed] [Google Scholar]
  184. St-Pierre MV, Serrano MA, Macias RI, Dubs U, Hoechli M, Lauper U, Meier PJ, and Marin JJ 2000. Expression of members of the multidrug resistance protein family in human term placenta. Am. J. Physiol. Regul. Integr. Comp. Physiol 279:R1495–R1503. [DOI] [PubMed] [Google Scholar]
  185. Sugawara N, Lai YR, Sugaware C, and Arizono K 1998. Decreased hepatobiliary secretion of inorganic mercury, its deposition and toxicity in the Eisai hyperbilirubinemic rat with no hepatic canalicular organic anion transporter. Toxicology 126:23–31. [DOI] [PubMed] [Google Scholar]
  186. Swensson A, and Ulfvarson U 1968. Distribution and excretion of various mercury compounds after single injections in poultry. Acta Pharmacol. Toxicol. Copenh 26:259–272. [DOI] [PubMed] [Google Scholar]
  187. Tanaka-Kagawa T, Naganuma A, and Imura N 1993. Tubular secretion and reabsorption of mercury compounds in mouse kidney. J. Pharmacol. Exp. Ther 264:776–782. [PubMed] [Google Scholar]
  188. Tanaka T, Naganuma A, and Imura N 1990. Role of gamma-glutamyltranspeptidase in renal uptake and toxicity of inorganic mercury in mice. Toxicology 60:187–198. [DOI] [PubMed] [Google Scholar]
  189. Tanaka T, Naganuma A, Kobayashi K, and Imura N 1991. An explanation for strain and sex differences in renal uptake of methylmercury in mice. Toxicology 69:317–329. [DOI] [PubMed] [Google Scholar]
  190. Tanaka T, Naganuma A, Miura N, and Imura N 1992. Role of testosterone in gamma-glutamyltranspeptidase-dependent renal methylmercury uptake in mice. Toxicol. Appl. Pharmacol 112:58–63. [DOI] [PubMed] [Google Scholar]
  191. Taugner R 1966. [On the renal uptake and intrarenal distribution of sublimate and Hg-cysteine]. Arzneimittelforschung 16:1120–1121. [PubMed] [Google Scholar]
  192. Taugner R, Winkel K, and Iravani J 1966. [On the localization of mercuric chloride concentration in the rat kidney]. Virchows Arch. Pathol. Anat. Physiol. Klin. Med 340:369–383. [PubMed] [Google Scholar]
  193. Terlouw SA, Graeff C, Smeets PH, Fricker G, Russel FG, Masereeuw R, and Miller DS 2002. Short- and long-term influences of heavy metals on anionic drug efflux from renal proximal tubule. J. Pharmacol. Exp. Ther 301:578–585. [DOI] [PubMed] [Google Scholar]
  194. Thomas DJ, and Smith JC 1979. Partial characterization of a low-molecular-weight methylmercury complex in rat cerebrum. Toxicol. Appl. Pharmacol 47:547–556. [DOI] [PubMed] [Google Scholar]
  195. Thomas DJ, and Smith JC 1982. Effects of coadministered low-molecular-weight thiol compounds on short-term distribution of methyl mercury in the rat. Toxicol. Appl. Pharmacol 62:104–110. [DOI] [PubMed] [Google Scholar]
  196. Ullrich KJ, Rumrich G, Fritzsch G, and Kloss S 1987. Contraluminal para-aminohippurate PAH transport in the proximal tubule of the rat kidney. II. Specificity: Aliphatic dicarboxylic acids. Pflugers Arch. 408:38–45. [DOI] [PubMed] [Google Scholar]
  197. Urano T, Iwasaki A, Himeno S, Naganuma A, and Imura N 1990. Absorption of methylmercury compounds from rat intestine. Toxicol. Lett 50:159–164. [DOI] [PubMed] [Google Scholar]
  198. Wagner CA, Lang F, and Broer S 2001. Function and structure of heterodimeric amino acid transporters. Am. J. Physiol. Cell Physiol 281:C1077–C1093. [DOI] [PubMed] [Google Scholar]
  199. Wakita Y 1987. Hypertension induced by methyl mercury in rats. Toxicol. Appl. Pharmacol 89:144–147. [DOI] [PubMed] [Google Scholar]
  200. Wang W, Clarkson TW, and Ballatori N 2000. Gamma-glutamyl transpeptidase and l-cysteine regulate methylmercury uptake by HepG2 cells, a human hepatoma cell line. Toxicol. Appl. Pharmacol 168:72–78. [DOI] [PubMed] [Google Scholar]
  201. Warkany J, and Hubbard DM 1953. A crodynia and mercury. J. Pediatr 42:365–386. [DOI] [PubMed] [Google Scholar]
  202. Wei H, Qiu L, Divine KK, Ashbaugh MD, McIntyre LC Jr., Fernando Q, and Gandolfi AJ 1999. Toxicity and transport of three synthesized mercury–thiol complexes in isolated rabbit renal proximal tubule suspensions. Drug Chem Toxicol 22:323–341. [DOI] [PubMed] [Google Scholar]
  203. World Health Organization. 2007. Mercury.
  204. Wu G 1995. Screening of potential transport systems for methyl mercury uptake in rat erythrocytes at 5 degrees by use of inhibitors and substrates. Pharmacol Toxicol 77:169–176. [DOI] [PubMed] [Google Scholar]
  205. Wu G 1996. Methylmercury-cysteine uptake by rat erythrocytes: Evidence for several transport systems. J. Appl. Toxicol 16:77–83. [DOI] [PubMed] [Google Scholar]
  206. Wu G 1997. Effect of probenecid on the transport of methyl mercury in erythrocytes by the organic anion transport system. Arch. Toxicol 71:218–222. [DOI] [PubMed] [Google Scholar]
  207. Yasutake A, Hirayama K, and Inoue M 1989. Mechanism of urinary excretion of methylmercury in mice. Arch. Toxicol 63:479–483. [DOI] [PubMed] [Google Scholar]
  208. Yokooji T, Murakami T, Yumoto R, Nagai J, and Takano M 2007. Site-specific bidirectional efflux of 2,4-dinitrophenyl-S-glutathione, a substrate of multidrug resistance-associated proteins, in rat intestine and Caco-2 cells. J. Pharm. Pharmacol 59:513–520. [DOI] [PubMed] [Google Scholar]
  209. Zalups RK 1991a. Autometallographic localization of inorganic mercury in the kidneys of rats: Effect of unilateral nephrectomy and compensatory renal growth. Exp. Mol. Pathol 54:10–21. [DOI] [PubMed] [Google Scholar]
  210. Zalups RK 1991b. Method for studying the in vivo accumulation of inorganic mercury in segments of the nephron in the kidneys of rats treated with mercuric chloride. J. Pharmacol. Methods 26:89–104. [DOI] [PubMed] [Google Scholar]
  211. Zalups RK 1991c. Renal accumulation and intrarenal distribution of inorganic mercury in the rabbit: effect of unilateral nephrectomy and dose of mercuric chloride. J. Toxicol. Environ. Health 33:213–228. [DOI] [PubMed] [Google Scholar]
  212. Zalups RK 1993a. Early aspects of the intrarenal distribution of mercury after the intravenous administration of mercuric chloride. Toxicology 79:215–228. [DOI] [PubMed] [Google Scholar]
  213. Zalups RK 1993b. Influence of 2,3-dimercaptopropane-1-sulfonate DMPS and meso-2,3-dimercaptosuccinic acid DMSA on the renal disposition of mercury in normal and uninephrectomized rats exposed to inorganic mercury. J. Pharmacol. Exp. Ther 267:791–800. [PubMed] [Google Scholar]
  214. Zalups RK 1995. Organic anion transport and action of gamma-glutamyl transpeptidase in kidney linked mechanistically to renal tubular uptake of inorganic mercury. Toxicol. Appl. Pharmacol 132:289–298. [DOI] [PubMed] [Google Scholar]
  215. Zalups RK 1997. Enhanced renal outer medullary uptake of mercury associated with uninephrectomy: implication of a luminal mechanism. J. Toxicol. Environ. Health 50:173–194. [DOI] [PubMed] [Google Scholar]
  216. Zalups RK 1998a. Basolateral uptake of inorganic mercury in the kidney. Toxicol. Appl. Pharmacol 151:192–199. [DOI] [PubMed] [Google Scholar]
  217. Zalups RK 1998b. Basolateral uptake of mercuric conjugates of N-acetylcysteine and cysteine in the kidney involves the organic anion transport system. J. Toxicol. Environ. Health A 55:13–29. [DOI] [PubMed] [Google Scholar]
  218. Zalups RK 1998c. Intestinal handling of mercury in the rat: implications of intestinal secretion of inorganic mercury following biliary ligation or cannulation. J. Toxicol. Environ. Health A 53:615–636. [DOI] [PubMed] [Google Scholar]
  219. Zalups RK 2000. Molecular interactions with mercury in the kidney. Pharmacol. Rev 52:113–143. [PubMed] [Google Scholar]
  220. Zalups RK, and Ahmad S 2004. Homocysteine and the renal epithelial transport and toxicity of inorganic mercury: Role of basolateral transporter organic anion transporter 1. J. Am. Soc. Nephrol 15:2023–2031. [DOI] [PubMed] [Google Scholar]
  221. Zalups RK, and Ahmad S 2005a. Handling of cysteine S-conjugates of methylmercury in MDCK cells expressing human OAT1. Kidney Int 68:1684–1699. [DOI] [PubMed] [Google Scholar]
  222. Zalups RK, and Ahmad S 2005b. Handling of the homocysteine S-conjugate of methylmercury by renal epithelial cells: Role of organic anion transporter 1 and amino acid transporters. J. Pharmacol. Exp. Ther 315:896–904. [DOI] [PubMed] [Google Scholar]
  223. Zalups RK, and Ahmad S 2005c. Transport of N-acetylcysteine s-conjugates of methylmercury in Madin-Darby canine kidney cells stably transfected with human isoform of organic anion transporter 1. J. Pharmacol. Exp. Ther 314:1158–1168. [DOI] [PubMed] [Google Scholar]
  224. Zalups RK, Aslamkhan AG, and Ahmad S 2004. Human organic anion transporter 1 mediates cellular uptake of cysteine-S conjugates of inorganic mercury. Kidney Int. 66:251–261. [DOI] [PubMed] [Google Scholar]
  225. Zalups RK, and Barfuss D 1990. Accumulation of inorganic mercury along the renal proximal tubule of the rabbit. Toxicol. Appl. Pharmacol 106:245–253. [DOI] [PubMed] [Google Scholar]
  226. Zalups RK, and Barfuss D 1993. Transport and toxicity of methylmercury along the proximal tubule of the rabbit. Toxicol. Appl. Pharmacol 121:176–185. [DOI] [PubMed] [Google Scholar]
  227. Zalups RK, and Barfuss D 1995. Pretreatment with p-aminohippurate inhibits the renal uptake and accumulation of injected inorganic mercury in the rat. Toxicology 103:23–35. [DOI] [PubMed] [Google Scholar]
  228. Zalups RK, and Barfuss D 1998a. Participation of mercuric conjugates of cysteine, homocysteine, and N-acetylcysteine in mechanisms involved in the renal tubular uptake of inorganic mercury. J. Am. Soc. Nephrol 9:551–561. [DOI] [PubMed] [Google Scholar]
  229. Zalups RK, and Barfuss D 1998b. Small aliphatic dicarboxylic acids inhibit renal uptake of administered mercury. Toxicol. Appl. Pharmacol 148:183–193. [DOI] [PubMed] [Google Scholar]
  230. Zalups RK, Barfuss DW, and Lash LH 1999a. Disposition of inorganic mercury following biliary obstruction and chemically induced glutathione depletion: Dispositional changes one hour after the intravenous administration of mercuric chloride. Toxicol. Appl. Pharmacol 154:135–144. [DOI] [PubMed] [Google Scholar]
  231. Zalups RK, Barfuss DW, and Lash LH 1999b. Relationships between alterations in glutathione metabolism and the disposition of inorganic mercury in rats: Effects of biliary ligation and chemically induced modulation of glutathione status. Chem. Biol. Interact 123:171–195. [DOI] [PubMed] [Google Scholar]
  232. Zalups RK, and Bridges CC 2009. MRP2 involvement in renal proximal tubular elimination of methylmercury mediated by DMPS or DMSA. Toxicol. Appl. Pharmacol 235:10–17. [DOI] [PubMed] [Google Scholar]
  233. Zalups RK and Diamond GL 1987a. Intrarenal distribution of mercury in the rat: Effect of administered dose of mercuric chloride. Bull. Environ. Contam. Toxicol 38:67–72. [DOI] [PubMed] [Google Scholar]
  234. Zalups RK and Diamond GL 1987b. Mercuric chloride-induced nephrotoxicity in the rat following unilateral nephrectomy and compensatory renal growth. Virchows Arch. B Cell. Pathol. Incl. Mol. Pathol 53:336–346. [DOI] [PubMed] [Google Scholar]
  235. Zalups RK, and Lash LH 1994. Advances in understanding the renal transport and toxicity of mercury. J. Toxicol. Environ. Health 42:1–44. [DOI] [PubMed] [Google Scholar]
  236. Zalups RK, and Lash LH 1997a. Binding of mercury in renal brush-border and basolateral membrane-vesicles. Biochem. Pharmacol 53:1889–1900. [DOI] [PubMed] [Google Scholar]
  237. Zalups RK, and Lash LH 1997b. Depletion of glutathione in the kidney and the renal disposition of administered inorganic mercury. Drug Metab. Dispos 25:516–523. [PubMed] [Google Scholar]
  238. Zalups RK, and Minor KH 1995. Luminal and basolateral mechanisms involved in the renal tubular uptake of inorganic mercury. J. Toxicol. Environ. Health 46:73–100. [DOI] [PubMed] [Google Scholar]

RESOURCES