Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2017 Dec 1.
Published in final edited form as: J Trace Elem Med Biol. 2016 Mar 4;38:99–107. doi: 10.1016/j.jtemb.2016.03.001

Methylmercury and brain development: A review of recent literature

Alessandra Antunes dos Santos 1,*, Mariana Appel Hort 2, Megan Culbreth 1, Caridad López-Granero 1, Marcelo Farina 3, Joao BT Rocha 4, Michael Aschner 1,*
PMCID: PMC5011031  NIHMSID: NIHMS769156  PMID: 26987277

Abstract

Methylmercury (MeHg) is a potent environmental pollutant, which elicits significant toxicity in humans. The central nervous system CNS) is the primary target of toxicity, and is particularly vulnerable during development. Maternal exposure to MeHg via consumption of fish and seafood can have irreversible effects on the neurobehavioral development of children, even in the absence of symptoms in the mother. It is well documented that developmental MeHg exposure may lead to neurological alterations, including cognitive and motor dysfunction. The neurotoxic effects of MeHg on the developing brain have been extensively studied. The mechanism of toxicity, however, is not fully understood. No single process can explain the multitude of effects observed in MeHg-induced neurotoxicity. This review summarizes the most current knowledge on the effects of MeHg during nervous system development considering both, in vitro and in vivo experimental models. Considerable attention was directed towards the role of glutamate and calcium dyshomeostasis, mitochondrial dysfunction, as well as the effects of MeHg on cytoskeletal components/regulators.

Keywords: Methylmercury, developmental neurotoxicity, mechanisms

1. Introduction

Mercury (Hg) is a heavy metal present in the environment that originates from from both natural and anthropogenic sources [1]. It occurs in nature as both inorganic (elemental or Hg0, Hg+ or Hg2+) and organic compounds, such as ethylmercury (EtHg) and methylmercury (MeHg) [2]. MeHg is produced by biomethylation of inorganic mercury present in aquatic sediments, a reaction catalyzed primarily by aquatic microorganisms [3, 4]. It accumulates up the aquatic food chain, and reaches maximal concentrations in long-lived, predatory fish such as tuna, swordfish, shark and whale [1, 3]. As a consequence, populations whose diet consists largely of fish and seafood may be potentially exposed to high levels of MeHg, thus rendering these communities vulnerable to toxicity [2, 5].

MeHg has an extensive toxicological history, with adverse effects in several organ systems observed throughout human life span [1, 6]. As mentioned above, seafood represents a major source of MeHg, and about 95% of that ingested is absorbed in the gastrointestinal tract [7]. After absorption, MeHg is ubiquitously distributed, and readily penetrates the CNS [8]. The brain has high affinity for MeHg, and concentrations can be 3–6 times greater than that in blood [6]. MeHg can distribute to all brain regions by crossing the blood-brain barrier via the neutral amino acid transport system L as a complex with L-cysteine [9].

It is well documented that the developing CNS is particularly vulnerable to MeHg [1012]. Taking that into account, this review summarizes the current knowledge regarding the molecular mechanisms involved in MeHg-induced developmental neurotoxicity. Considerable attention has been directed towards the role of glutamate and calcium dyshomeostasis, mitochondrial dysfunction as well as the effects of MeHg on cytoskeletal components/regulators.

1.1. Neurotoxic Effects of Methylmercury – General Aspects

The CNS is a major target of MeHg [13]. In adults (i.e. the mature CNS) toxic effects of MeHg are characterized by a long latency period before the appearance of symptoms [3]. The initial symptoms include blurred vision, weight loss, paresthesias of the circumoral area and hands and feet, followed by visual-field constriction, ataxia and psychiatric sympotmatology [3, 14]. Furthermore, in adults the signs and symptoms of MeHg poisoning are associated with loss of neuronal cells in specific brain regions, such as the visual cortex and the cerebellum [15]. Post-mortem histopathological analysis has shown a significant loss of cerebellar granule cells (CGC) with relative preservation of the adjacent Purkinje cell layer [16].

The relationship between MeHg-induced motor deficit and cerebellar damage is, in fact, a well-described phenomenon [17]. Several studies have shown that exposure to MeHg can produce behavioral deficits related to locomotor activity and motor performance in adult mice [1821]. Experimental studies conducted in the last four decades have contributed to the current understanding of pivotal events that mediate MeHg-induced neurotoxicity. Disruption of calcium (Ca2+) homeostasis or induction of oxidative stress via overproduction of reactive oxygen species (ROS) or reduction of antioxidant defenses are likely to be critical factors in MeHg-induced cell damage, as are interactions with sulfhydryl (-SH) groups [5, 2225]. The high affinity of MeHg for -SH groups on the amino acids cysteine and methionine may alter the structure of a large number of proteins and lead to disruption of various intracellular functions. A number of molecular targets and mechanisms have been implicated in both in vitro and in vivo studies and no single mechanism is likely to be explanatory [22, 26, 27].

In the CNS, MeHg can adversely affect many cell types, and a large body of evidence suggests that astrocytes represent a preferential site for its accumulation, playing a crucial role in MeHg neurotoxicity [2830]. Astrocytes have an essential role in brain development, function, and plasticity. These cells coordinate neuronal development and survival, control synapse formation and function, and play a major role in the formation of neuronal circuits [3134]. Moreover, astrocytes carry out other critical functions in the brain, including formation of the blood brain barrier and expression of neurotransmitter transporters [35]. MeHg interferes with glutamate and aspartate uptake in astrocytes, increasing glutamate concentration in the synaptic cleft, resulting in high levels of extracellular glutamate, which is toxic to neurons [25, 3638]. Furthermore, MeHg inhibits the astrocytic uptake of cystine and cysteine, compromising glutathione (GSH) synthesis and cellular redox status [3941]. In addition to its effects in astrocytes, MeHg can also stimulate the production and secretion of lysosomal proteases in microglial cells, also leading to neuronal toxicity [42].

1.2. Effects of MeHg on the Developing Brain

Prolonged pre and/or perinatal exposure to MeHg, even at moderate doses, is related to multiple deficits in neurons and glia including abnormal migration, differentiation and growth [43, 44]. The developing brain is extremely sensitive to MeHg poisoning. In fact, children exposed to MeHg may exhibit decreased IQ, impaired movements, visuospatial perception, and speech, a medical phenomenon known as fetal Minamata disease (FMD) [4547]. The effects of developmental exposure to MeHg on animal behavior include, among others, reduced motor activity [52] and a decrease in memory [5355] and learning [56]. Studies modeling in utero human exposures in postnatal rat pups, show that MeHg can alter the expression of many developmental regulators and genes involved in small GTPase signaling pathways regulating cell growth and proliferation [57]. Furthermore, several reports have shown that MeHg can induce mitotic arrest and caspase-dependent apoptosis in the developing CNS [27, 5860], suggesting that proliferative cell populations may be especially vulnerable to MeHg-induced cell death.

In 1952, the first case of congenital MeHg poisoning was described in Sweden. Children from a family that used flour made from a MeHg-contaminated seed grain, were found to have impaired intellectual function and motor development [61]. Some years later, a catastrophic epidemic from environmental MeHg contamination was reported in Minamata, Japan. The ingestion of contaminated fish by pregnant women evidently poisoned the fetuses, and consequently many children born from 1955 onward had neurodevelopmental disturbances [62]. More recently, studies conducted in human populations found an association between maternal MeHg exposure during pregnancy, and neurological and neuropsychological deficits in children [6365].

Characteristics and mechanisms of MeHg poisoning cases in adults are distinct from that in children [15]. In the neonatal period or at early ages, the damage caused by MeHg is more widely distributed throughout the CNS [3, 12, 66]. The injury is pervasive, and often, the earlier the exposure, the more generalized the damage observed [66]. Experimental data have provided evidence that MeHg can affect processes that are extremely important to brain development, such as: (i) interaction with elements of the cytoskeleton that consequently affect the movement of neuronal cells from the site of germination to the final site of function [67, 68] (ii) interference with the formation of dendrites and axons (i.e. interneuronal contacts) [68] and (iii) induction of apoptosis, which occurs normally during the development of the brain [69, 70].

The next section provides an overview of recent findings regarding MeHg-induced developmental neurotoxicity, as well as a discussion about the potential molecular mechanisms of MeHg-induced cell death that have recently been described in the literature using animal and cell culture models.

2. Mechanisms of MeHg-induced Developmental Neurotoxicity

2.1. Glutamate and Ca2+

It has been demonstated that MeHg-induced neuronal death can occur either by necrosis [7173] or apoptosis [60, 70, 71, 7376], depending on the level of exposure, the cell type, and the cellular defense mechanisms. In many instances, such effects are mediated by disruption/modulation of cellular Ca2+ homeostasis or increased generation of ROS in the mitochondria or other intracellular compartments [77]. In fact, Ca2+ is known to play a critical role in cell loss in the CNS and Ca2+ overload has been shown to trigger either necrotic or apoptotic cell death [for review, see Orrenius et al., 2003 - [78]. Sustained elevated Ca2+ levels have been found in various cell types after MeHg exposure, and protective effects of Ca2+-chelators or Ca2+-channel blockers have been reported in vitro and in vivo [7984]. Importantly, MeHg at low micromolar concentrations causes prolonged increases in intracellular, cytosolic Ca2+ concentrations [79, 80, 85, 86], which may be secondary to increases in extracellular glutamate [25]. A number of experimental findings point to inhibition of glutamate uptake by astrocytes, increase in spontaneous release of glutamate from presynaptic terminals, and inhibition of vesicular glutamate uptake as critical phenomena linked to MeHg-mediated excitotoxicity [25, 41, 87]. During critical CNS developmental windows, MeHg-induced disruption of the glutamate pathway may interfere with cell proliferation and cell fate decisions. Using a MeHg drinking water exposure paradigm in mouse dams, Manfroi et al. [88] investigated the exclusive contribution of MeHg exposure through maternal milk. This report demonstrated that the offspring of the exposed dams exhibitied glutamate uptake inhibition in the cerebellum; this was not observed in the dams, however. Experimental data obtained in cortical cell cultures from pups exposed to MeHg during gestation (8 mg/kg, gestational-days 8 or 15), showed effects on glutamatergic neurotransmission and increased responsiveness of the N-methyl-D-aspartate (NMDA)-type glutamate receptors. The basal extracellular glutamate levels measured in these cultures was higher than those measured in cultures obtained from control pups [53]. This suggests an enhanced sensitivity of NMDA receptors, which could make premature cortical neurons more susceptible to MeHg neurotoxicity. Interestingly, MK-801, a non-competitive antagonist of NMDA receptors, administered intraperitoneally with MeHg during brain development, markedly ameliorated MeHg-induced neurodegeneration [89]. All the above mentioned observations are consistent with MeHg increasing extracellular glutamate levels. Overactivation of NMDA receptors increases Ca2+ influx into neurons, therefore leading to activation of important pathways involved with cell death. Alternatively, Ca2+ can be taken up by mitochondria, where it may stimulate ROS production [90]. In this regard, MeHg induces pro-oxidative damage in the developing CNS, by disrupting postnatal development of the glutathione-antioxidant system [91]. Furthermore, it has been shown that exclusive lactational exposure to MeHg causes neurotoxicity in pups by decreasing glutamate uptake into cerebellar slices and increasing oxidative stress [88, 92].

Neurons in patients with FMD are hypoplastic, ectopic, and disoriented, indicating disrupted migration, maturation, and growth. As previously mentioned, the cerebellum is one of the most susceptible brain regions to MeHg exposure, and profound loss of cerebellar granule cells (CGCs) is detected in the brains of patients with FMD [93]. MeHg causes a characteristic lack of CGC migration, and subsequently laminar cortical organization in the developing cerebellum in vivo [66, 93]. Significant impairment of CGC migration by MeHg has been shown in vitro [94] and in organotypic slice culture under acute and/or high level [67] or repeated exposure to low micromolar concentrations, as well [95]. The question of which signaling molecules are involved in these effects on neuronal cell migration, however, remains to be determined. Among the early effects of MeHg in CGCs is a pronounced increase of Ca2+, consisting of an initial release of stored intracellular Ca2+ followed by influx of extracelular Ca2+ [79, 96]. This effect occurs at much lower MeHg concentrations in CGCs than in other, less sensitive cells [97, 98]. It has been shown that MeHg-induced neuronal death is attenuated by chelating Ca2+ in vitro [80], while in vivo, the neurological signs of MeHg toxicity in rats are attenuated by Ca2+ channel blockers [82], suggesting that loss of Ca2+ homeostasis is critical to MeHg-induced death of CGCs. Nevertheless, a study conducted in CGCs exposed to nanomolar concentrations of MeHg (300 nM) for extended periods of time (4 days in vitro), revealed that the Ca2+ concentration remained unchanged, while a significant reduction of glutathione peroxidase-1 (GPx-1) activity was observed [99]. GPx-1 is a crucial antioxidant enzyme involved in neuronal detoxification of hydrogen peroxide [100]. Interestingly, the authors showed that GPx-1 inhibition occurred before any changes on potential targets of MeHg toxicity, and GPx1 overexpression was able to prevent MeHg-induced neuronal death in CGCs [99]. Recently, it has been proposed that alterations of Ca2+ spike frequency could be one of the primary targets of MeHg-induced inhibition of granule cell migration, in a mouse model of FMD. The authors showed that MeHg slows the speed of granule cell migration and reduces the frequency of spontaneous Ca2+ spikes in granule cell somata in a dose-dependent manner [93]. Ca2+ spikes (or transient Ca2+ elevations) are known to play an essential role in maintaining the movement of cells ranging from fibroblasts to immature neurons [101103]. Thus, the inhibition of Ca2+ spikes by buffering of intracellular Ca2+ levels reduces or inhibits cell movement [104].

Alterations in voltage-sensitive Ca2+ channels (VSCCs) have also been reported in MeHg-induced neurotoxicity. It is known that Ca2+ channels are expressed early during development in many parts of the CNS. These channels participate in many intracellular events such as neurotransmitter release, axonal growth, and gene expression by allowing influx of extracellular Ca2+, resulting in the subsequent activation of proteins [105, 106]. Neonatal CGCs in culture possess a rich diversity of Ca2+ channels, with five postulated types (L, N, P, Q, and R) being described to date [107]. Migration of CGCs, which is inhibited by MeHg exposure, is dependent on N-type VSCCs [108, 109]. It has been reported that Ca2+ channels are a likely MeHg target [81, 110, 111] because of their location on the plasma membrane and strict regulation by various intracellular mediators [81]. Acute and chronic effects of MeHg on VSCCs have been characterized in PC12 cells [110112], a well characterized model for differentiated neurons. In neonatal CGCs, a more sensitive cell type, this toxicant can inhibit VSCC function even at submicromolar concentrations [81].

A recent study reported that acute exposure to MeHg causes developmental stage-dependent increases in Ca2+ concentration in slices of neonatal rat cerebellum. The effects of MeHg were most prominent in CGCs during development or early stages of migration. In addition, the authors showed that the effects of MeHg on CGCs during migration is attenuated by gamma-aminobutyric acid (GABAA) receptor modulation, indicating that this receptor can be affected in any stage of CGC development [113]. In fact, studies conducted in hippocampal and cerebellar slice preparations revealed that MeHg can impair inhibitory GABAergic neurons [114116]. Moreover, when compared with glutamatergic synaptic transmission, hippocampal GABAergic transmission seems to be more sensitive to the effects of MeHg [116]. The effect of several developmental neurotoxicants on mRNA levels of various neural markers was analyzed in CGCs, and results revealed that prolonged exposure to 50 nM MeHg resulted in a significant downregulation of GABAA receptor mRNA, while other neural markers were unchanged [117].

2.2. Mitochondria

Mitochondria represent the major sites for oxidative stress, and are critical for regulation of neuronal cell death induced by MeHg via Ca2+ dyshomeostasis and/or ROS generation [7880]. Indeed, this organelle has been considered the main source of intracellular Ca2+ release in CGC cultures upon MeHg exposure [96]. Previous studies suggest that MeHg acts via a mitochondrial-dependent cascade increasing Bax, cytosolic cytochrome c, and caspase-9, to elicit cell death in primary neuronal precursors and cell lines [76]. Acute MeHg exposure was able to increase the levels of activated caspase-3 in the postnatal rat hippocampus, and cause subsequent juvenile learning deficits [59, 27]. Furthermore, exposure of the developing hippocampus in vivo induced mitochondrial-dependent cell death via sequential activation of caspase-9 and caspase-3 [60]. In fact, it is well known that apoptotic cell death occurs via activation of well characterized biochemical pathways, with caspase-8 and caspase-9 acting as initiator caspases that cleave and activate caspase-3, which is normally considered one of the last steps in cell death [118]. In differentiating PC12 cells, immunocytochemical analysis showed that activated caspase-3-positive neurons were significantly increased 1 day after exposure to 100 nM MeHg. Importantly, it was also observed apoptosis preceded inhibition of neurite extension [119]. In the developing brain, caspase-3 has been found to play a crucial role not only in apoptosis [120], but also in nonapoptotic neuronal functions, including neural stem cell differentiation [121].

In rats, peak of postnatal neurogenesis (postnatal day 7– P7) corresponds to human hippocampal development occurring during the third trimester of gestation. Deficits in hippocampal-dependent behavioral tasks might be a sign of impaired neurogenesis [122]. MeHg exposure in P7 rats led to reductions in cell cycle regulator cyclin E, as well as induction of mitochondrial-dependent apoptosis by activation of caspase-3 [59, 60]. Rats exposed to a single injection of 5μg/g MeHg at P7 showed profound juvenile spatial learning impairment and decreases in cell number in P21 hippocampus, that correlated with acute caspase-dependent apoptosis [27]. Furthermore, acute exposure to an environmentally relevant dose of MeHg (0.6μg/g, 24h) in P7 rats, induced caspase-3 activation in neural stem cells (NSCs) of the dentate gyrus (DG), and caused hippocampal-dependent memory deficits during adolescence, assessed by Morris water maze [123]. In addition, a significant degeneration of neuritic processes, and either apoptotic or necrotic death was observed in cortical cultures obtained from pups born to 8 mg/kg MeHg-exposed dams, which was correlated with long term memory impairment of adult offspring subjected to a memory task [73].

While pro-apoptotic signals stimulated by MeHg may involve caspase-3 activation, the overall process is likely to be more complex. For example, ROS-induced mitochondrial dysfunction causes the release of Ca2+ into the cytosol, which can trigger the activation of calpains, a calcium-dependent protease. A number of in vitro studies performed in CGCs and hippocampal cells after MeHg treatment reported mitochondrial dysfunction (determined by cytochrome C release) and Ca2+ influx in association with increased calpain protein level [124126]. Calpains along with caspases belong to the family of cysteine proteases, which play important roles in various necrotic and apoptotic conditions [127]. This pathway is activated by the perturbation of intracelular Ca2+ homeostasis[128], which can occur during exposure to various toxic stimuli, such as MeHg. Calpains are known to regulate the activity of several protein kinases and phosphatases that modify the cytoskeleton, in addition to cleaving diverse cellular substrates including cytoskeletal proteins directly [129, 130], as well as tubulins. Very low levels of MeHg (30nM) can induce apoptotic cell death of cultured rat cerebellar neurons via increased intracellular Ca2+ concentration and calpain activation [125]. It has been shown that calpains are activated in MeHg-exposed NSCs [76], with an associated increase in intracelular Ca2+ as observed in other apoptotic models [131, 132]. Interestingly, MeHg-exposed NSCs undergoing apoptosis showed activation of two parallel pathways involving caspases and calpains, as proven by either the partial protection exerted by the caspase (zVAD-fmk) or calpain (E64d) inhibitor alone, or the full protective effect of the two inhibitors together [76].

2.3. Cytoskeleton

The architecture and survival of neurons depends significantly upon their cytoskeletal components, specifically actin filaments and microtubules, such as β-tubulin. Microtubules are involved in a variety of cell functions important for the development and maintenance of cell shape, reproduction and division, signaling, intracellular transport, and motiliy. In this way, any disturbance of these dynamic structures or their polymerization/depolymerization status could compromise cell survival [133, 134]. MeHg has long been known to be a potent inhibitor of microtubule polymerization [135137]. In vitro studies demonstrated that MeHg has high affinity for tubulin sulphydryl groups (-SH), depolymerizing cerebral microtubules, and directly inhibiting their assembly [135, 137]. Microtubule fragmentation and neuronal network dissolution have been observed in cultured primary cerebellar granule neurons exposed to concentrations of MeHg as low as 0.5–1 μM. Such effects occurred early, and preceded the onset of apoptotic signs in this in vitro model [71]. More recently, a proteomic study of CGCs exposed to MeHg (≤ 1μM) for several days revealed that MeHg can disrupt the balance between phosphorylated and non-phosphorylated cofilin forms [138]. The balance of cofilin phosphorylation/dephosphorylation is a central component of actin dynamics in migrating cells and facilitates actin filament turnover, which is responsible for neuron shape, migration, polarity formation, and regulation of synaptic structure and function [139].

The Rho/ROCK signaling pathway, one of the most well-studied Rho family GTPase-effector pathways, is the major regulator of stress fiber formation under most physiological conditions. This pathway is closely related to the pathogenesis of several CNS disorders, and involved in many aspects of neuronal function, including neurite outgrowth and retraction [140]. Among Rho family proteins Ras-related C3 botulinum toxin substrate 1 (Rac1) and cell division cycle 42 (Cdc42) are known to promote neurite extension and inhibit apoptosis [141]. Moreover, the Rho-associated protein kinases (ROCKs) are central regulators of the actin cytoskeleton downstream of the small GTPase Rho [142]. Along this line, a study reported downregulation of Rho-family proteins Rac1 and Cdc42 in early stages of MeHg-induced cytotoxicity. The authors demonstrated that neuritic degeneration precedes MeHg-induced apoptotic cell death in cortical neurons [143]. The therapeutic effects of Rho/ROCK inhibitors have been demonstrated in some in vitro and in vivo experimental models, and this pathway has become an attractive target for the development of drugs for treating CNS disorders [144]. In this regard, treatment of cultured cortical neuronal cells with Rho/ROCK inhibitors suppressed MeHg-induced neuronal degeneration and apoptotic cell death at 100 nM MeHg for 3 days. Furthermore, the Rho/ROCK inhibitors partially prevented MeHg-induced toxicity in a rat model (20 ppm MeHg in drinking water for 28 days) [145].

2.4. Cell Differentiation

Normal differentiation of neurons and glial cells during development is crucial in neurogenesis, which is particularly sensitive to environmental toxicants like MeHg. In this regard, in vitro cultures of NSCs provide an appropriate model for studying the adverse effects of MeHg and other developmental neurotoxicants on survival, proliferation, differentiation, and migration. NSCs are self-renewing cells, that play an essential role in the development and maturation of the CNS, with potential to differentiate into neurons, astrocytes, and oligodendrocytes [146]. Nanomolar (2.5–10 nM) concentrations of MeHg can affect proliferation and differentiation of rodent embryonic NSCs, and the activation of Notch signaling seems to be involved in this process [147]. In fact, constitutive expression of Notch protein is reported to inhibit neurite outgrowth in PC12 cells [148]. Moreover, it has been shown that low concentrations of MeHg (2.5–25 nM) reduce neural progenitor cell proliferation [76, 149151], expression of genes related to cell cycle regulation (cyclin E, p16 and p21) [149, 151, 140], cellular senescence, and mitochondrial function (repression of the mitochondrial respiratory chain enzymes complexes I and III) [149]. Interestingly, the alterations detected in cell proliferation, cell cycle regulation (p16 and p21 regulators), and mitochondrial function, were observed in cells directly exposed to MeHg and in their daughter cells (from passages 2 and 3) cultured under MeHg-free conditions, suggesting that the effects of MeHg on NSCs are heritable [149].

3. Conclusion

MeHg is a hazardous environmental pollutant, of great concern to public health because of its toxicity to the CNS. Exposure events have demonstrated that fetuses are much more sensitive to MeHg than adults. It has been reported that acute or chronic MeHg exposure can cause adverse effects during all developmental periods. Indeed, MeHg is now recognized as an important developmental neurotoxicant, and over the last decade, notable progress has been made in understanding its mechanism of neurotoxicity. One of the most widely documented effects caused by MeHg on the CNS is associated with glutamate-mediated excitotoxicity, which can be linked to or followed by intracellular Ca2+ overload. In fact, impairment of intracellular calcium and glutamate homeostasis, oxidative stress generation, as well as mitochondrial dysfunction are critical factors in MeHg-induced cell death. These connected phenomena create complexity in understanding the initial events of neurotoxicity. Current knowledge of the precise targets of MeHg is incomplete, and detailed experimental data regarding the mechanism of action, mainly during the neurodevelopmental period, are needed. In this review, the mechanisms implicated in the published literature are particularly pertinent to MeHg-induced developmental neurotoxicity. For instance, MeHg-induced negative effects on cytoskeletal proteins (microtubules) and cytoskeleton-regulating proteins (Rho-family proteins), thereby disturbing neuronal migration and differentiation. Relatively little new information, however, has been found about the potentially neurotoxic effects induced by interaction between MeHg and cytoskeletal proteins. In addition, the relationship between the Rho/ROCK pathway and MeHg neurotoxicity is still very limited, and consideration should be given to future research. This review has provided an overview of the main molecular mechanims involved in neurotoxicity induced by MeHg-exposure during the developmental period. Taken together, the data discussed in this paper collaborate for a better understanding of the multifactorial mechanisms involved in MeHg-induced cell damage, suggesting, a number of directions for future research.

Table 1.

In vivo studies on molecular mechanisms of MeHg-induced developmental neurotoxicity

Experimental model Treatment Tissue Results Reference
Swiss albino mice Drinking water - 15 mg/l for 21 days Cerebellum of suckling mice indirectly exposed to MeHg for 21 days
  • Inhibition of glutamate uptake

  • Increased levels of hydroperoxide

[88]
Sprague–Dawley rats Acute - intragastric intubation (8 mg/kg on GD 8 and 15) Cortical cell cultures from pups exposed to MeHg during gestation
  • Increased extracellular glutamate levels

  • Increased responsiveness of the NMDA receptors

[53]
Swiss albino mice Drinking water - (1, 3 and 10 mg/l) during the gestational period Whole brain of P1, P11 and P21 pups
  • Dose-dependent inhibition of GSH levels

  • Dose-dependent inhibition of Gpx and GR activities

  • Increased F2-isoprostanes levels

[91]
Sprague–Dawley rat pups (P7) Acute - subcutaneous injection (5 μg/g) Hippocampus
  • Increased levels of activated caspase-3 (24 h after injection)

  • Decreased cell number in P21 hippocampus

[27]
Sprague–Dawley rat pups (P7) Acute - subcutaneous injection (5 μg/g) Hippocampus
  • Increased cytochrome c release

  • Increased Bax protein

  • Caspase-9 and caspase-3 activation

[60]
Sprague–Dawley rat pups (P7) Acute - subcutaneous injection (0.6 μg/g) NSCs from dentate gyrus hilus and granule cell layer
  • Acute caspase activation

  • Decreased cell number (P21)

  • Reduced neural precursor proliferation (P21)

[123]

Abbreviations: GD (gestational days), P (postnatal days), NSCs (neural stem cells), Gpx (glutathione peroxidase), GR (glutathione reductase)

Table 2.

In vitro studies on molecular mechanisms of MeHg-induced neurotoxicity

Experimental model Exposure dose Results Reference

Primary cerebellar granule cells (Sprague–Dawley rats) 0.5 μM
  • Increased [Ca+2]i concentration

  • Mitochondrial membrane depolarization

[96]

Cerebellar slices from Sprague-Dawley rats (P8–P12) 10 – 20 μM
  • Increased [Ca+2]i concentration in CGCs

[113]

Mouse neuronal progenitor cell line (C17.2) 0.25 – 0.5 μM
  • Bax activation

  • Cytochrome c release

  • Caspase activation

  • Calpain activation

[76]
Primary embryonic cortical NSCs (Sprague–Dawley rats) 0.025 – 0.05 μM

PC12 (rat pheochromocytoma cell line) 0.1 μM
  • Increased TUNEL-positive apoptotic nuclei

  • Caspase 3 activation

  • Inhibition of neurite extension

[119]

HT22 cells (mouse hippocampal cell line) 4.0 μM
  • Decreased mitochondrial membrane potential and function

  • Reduced ATP levels

  • Cytochrome c release

  • Calpain activation

  • Lysosomal disruption

[126]

Primary embryonic cortical NPCs (Sprague–Dawley rat) 10 nM
  • Cyclin E downregulation

  • GSK-3β upregulation

  • Reduced NPCs proliferation

[151]

Primary embryonic cortical NSCs (Sprague–Dawley rats) 2.5 – 5.0 nM
  • Reduced NSC proliferation

  • Increased expression of p16 and p21 (cell cycle arrest)

  • Altered expression of mitochondrial genes

[149]

Primary embryonic cortical culture (Sprague–Dawley rats) 0.1 – 1 μM
  • Neuritic degeneration

  • Caspase-dependent apoptosis

  • Rac1 and Cdc42 downregulation

[143]

Abbreviations: P (postnatal days), NSCs (neural stem cells), CGCs (cerebellar granule cells), NPCs (neural progenitor cells).

Acknowledgments

This work was funded by NIEHS R01 ES007331 and NIEHS R01 ES020852,

Footnotes

COI: All authors disclose no financial and personal relationships with other people or organisations that could inappropriately influence (bias) of this work.

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

References

  • 1.NRC. Toxicological Effects of Methylmercury. Washington (DC): 2000. Toxicological Effects of Methylmercury. [Google Scholar]
  • 2.Clarks TW. The three modern faces of mercury. Environ Health Perspect. 2002;110(Suppl 1):11–23. doi: 10.1289/ehp.02110s111. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Clarkson TW, Magos L, Myers GJ. The toxicology of mercury--current exposures and clinical manifestations. N Engl J Med. 2003;349(18):1731–1737. doi: 10.1056/NEJMra022471. [DOI] [PubMed] [Google Scholar]
  • 4.Hintelmann H. Organomercurials. Their formation and pathways in the environment. Met Ions Life Sci. 2010;7:365–401. doi: 10.1039/BK9781847551771-00365. [DOI] [PubMed] [Google Scholar]
  • 5.Farina M, Aschner M, Rocha JB. Oxidative stress in MeHg-induced neurotoxicity. Toxicol Appl Pharmacol. 2011;256(3):405–417. doi: 10.1016/j.taap.2011.05.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Syversen T, Kaur P. The toxicology of mercury and its compounds. J Trace Elem Med Biol. 2012;26(4):215–226. doi: 10.1016/j.jtemb.2012.02.004. [DOI] [PubMed] [Google Scholar]
  • 7.Kershaw TG, Clarkson TW, Dhahir PH. The relationship between blood levels and dose of methylmercury in man. Arch Environ Health. 1980;35(1):28–36. doi: 10.1080/00039896.1980.10667458. [DOI] [PubMed] [Google Scholar]
  • 8.Zareba G, Cernichiari E, Hojo R, Nitt SM, Weiss B, Mumtaz MM, Jones DE, Clarkson TW. Thimerosal distribution and metabolism in neonatal mice: comparison with methyl mercury. J Appl Toxicol. 2007;27(5):511–518. doi: 10.1002/jat.1272. [DOI] [PubMed] [Google Scholar]
  • 9.Aschner M, Clarkson TW. Uptake of methylmercury in the rat brain: effects of amino acids. Brain Res. 1988;462(1):31–39. doi: 10.1016/0006-8993(88)90581-1. [DOI] [PubMed] [Google Scholar]
  • 10.Costa LG, Aschner M, Vitalone A, Syversen T, Soldin OP. Developmental neuropathology of environmental agents. Annu Rev Pharmacol Toxicol. 2004;44:87–110. doi: 10.1146/annurev.pharmtox.44.101802.121424. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Johansson C, Castoldi AF, Onishchenko N, Manzo L, Vahter M, Ceccatelli S. Neurobehavioural and molecular changes induced by methylmercury exposure during development. Neurotox Res. 2007;11(3–4):241–260. doi: 10.1007/BF03033570. [DOI] [PubMed] [Google Scholar]
  • 12.Grandjean P, Herz KT. Methylmercury and brain development: imprecision and underestimation of developmental neurotoxicity in humans. Mt Sinai J Med. 2011;78(1):107–118. doi: 10.1002/msj.20228. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Ni M, Li X, Rocha JB, Farina M, Aschner M. Glia and methylmercury neurotoxicity. J Toxicol Environ Health A. 2012;75(16–17):1091–1101. doi: 10.1080/15287394.2012.697840. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Bakir F, Damluji SF, Amin-Zaki L, Murtadha M, Khalidi A, al-Rawi NY, Tikriti S, Dahahir HI, Clarkson TW, Smith JC, Doherty RA. Methylmercury poisoning in Iraq. Science. 1973;181(4096):230–241. doi: 10.1126/science.181.4096.230. [DOI] [PubMed] [Google Scholar]
  • 15.Aschner M, Syversen T. Methylmercury: recent advances in the understanding of its neurotoxicity. Ther Drug Monit. 2005;27(3):278–283. doi: 10.1097/01.ftd.0000160275.85450.32. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Takeuchi T. Pathology of Minamata disease. With special reference to its pathogenesis. Acta Pathol Jpn. 1982;32(Suppl 1):73–99. [PubMed] [Google Scholar]
  • 17.Sakamoto M, Nakano A, Kajiwara Y, Naruse I, Fujisaki T. Effects of methyl mercury in postnatal developing rats. Environ Res. 1993;61(1):43–50. doi: 10.1006/enrs.1993.1048. [DOI] [PubMed] [Google Scholar]
  • 18.Farina M, Franco JL, Ribas CM, Meotti FC, Missau FC, Pizzolatti MG, Dafre AL, Santos AR. Protective effects of Polygala paniculata extract against methylmercury-induced neurotoxicity in mice. J Pharm Pharmacol. 2005;57(11):1503–1508. doi: 10.1211/jpp.57.11.0017. [DOI] [PubMed] [Google Scholar]
  • 19.Carvalho MC, Franco JL, Ghizoni H, Kobus K, Nazari EM, Rocha JB, Nogueira CW, Dafre AL, Muller YM, Farina M. Effects of 2,3-dimercapto-1-propanesulfonic acid (DMPS) on methylmercury-induced locomotor deficits and cerebellar toxicity in mice. Toxicology. 2007;239(3):195–203. doi: 10.1016/j.tox.2007.07.009. [DOI] [PubMed] [Google Scholar]
  • 20.Zimmermann LT, dos Santos DB, Colle D, dos Santos AA, Hort MA, Garcia SC, Bressan LP, Bohrer D, Farina M. Methionine stimulates motor impairment and cerebellar mercury deposition in methylmercury-exposed mice. J Toxicol Environ Health A. 2014;77(1–3):46–56. doi: 10.1080/15287394.2014.865582. [DOI] [PubMed] [Google Scholar]
  • 21.Kirkpatrick M, Benoit J, Everett W, Gibson J, Rist M, Fredette N. The effects of methylmercury exposure on behavior and biomarkers of oxidative stress in adult mice. Neurotoxicology. 2015;50:170–178. doi: 10.1016/j.neuro.2015.07.001. [DOI] [PubMed] [Google Scholar]
  • 22.Ceccatelli S, Dare E, Moors M. Methylmercury-induced neurotoxicity and apoptosis. Chem Biol Interact. 2010;188(2):301–308. doi: 10.1016/j.cbi.2010.04.007. [DOI] [PubMed] [Google Scholar]
  • 23.Farina M, Rocha JB, Aschner M. Mechanisms of methylmercury-induced neurotoxicity: evidence from experimental studies. Life Sci. 2011;89(15–16):555–563. doi: 10.1016/j.lfs.2011.05.019. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Giordano G, Costa LG. Developmental neurotoxicity: some old and new issues. ISRN Toxicol. 2012;2012:814795. doi: 10.5402/2012/814795. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Aschner M, Syversen T, Souza DO, Rocha JB, Farina M. Involvement of glutamate and reactive oxygen species in methylmercury neurotoxicity. Braz J Med Biol Res. 2007;40(3):285–291. doi: 10.1590/s0100-879x2007000300001. [DOI] [PubMed] [Google Scholar]
  • 26.Limke TL, Bearss JJ, Atchison WD. Acute exposure to methylmercury causes Ca2+ dysregulation and neuronal death in rat cerebellar granule cells through an M3 muscarinic receptor-linked pathway. Toxicol Sci. 2004;80(1):60–68. doi: 10.1093/toxsci/kfh131. [DOI] [PubMed] [Google Scholar]
  • 27.Falluel-Morel A, Sokolowski K, Sisti HM, Zhou X, Shors TJ, Dicicco-Bloom E. Developmental mercury exposure elicits acute hippocampal cell death, reductions in neurogenesis, and severe learning deficits during puberty. J Neurochem. 2007;103(5):1968–1981. doi: 10.1111/j.1471-4159.2007.04882.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Charleston JS, Bolender RP, Mottet NK, Body RL, Vahter ME, Burbacher TM. Increases in the number of reactive glia in the visual cortex of Macaca fascicularis following subclinical long-term methyl mercury exposure. Toxicol Appl Pharmacol. 1994;129(2):196–206. doi: 10.1006/taap.1994.1244. [DOI] [PubMed] [Google Scholar]
  • 29.Charleston JS, Body RL, Bolender RP, Mottet NK, Vahter ME, Burbacher TM. Changes in the number of astrocytes and microglia in the thalamus of the monkey Macaca fascicularis following long-term subclinical methylmercury exposure. Neurotoxicology. 1996;17(1):127–138. [PubMed] [Google Scholar]
  • 30.Aschner M, Du YL, Gannon M, Kimelberg HK. Methylmercury-induced alterations in excitatory amino acid transport in rat primary astrocyte cultures. Brain Res. 1993;602(2):181–186. doi: 10.1016/0006-8993(93)90680-l. [DOI] [PubMed] [Google Scholar]
  • 31.Booth GE, Kinrade EF, Hidalgo A. Glia maintain follower neuron survival during Drosophila CNS development. Development. 2000;127(2):237–244. doi: 10.1242/dev.127.2.237. [DOI] [PubMed] [Google Scholar]
  • 32.Ullian EM, Sapperstein SK, Christopherson KS, Barres BA. Control of synapse number by glia. Science. 2001;291(5504):657–661. doi: 10.1126/science.291.5504.657. [DOI] [PubMed] [Google Scholar]
  • 33.Clarke LE, Barres BA. Emerging roles of astrocytes in neural circuit development. Nat Rev Neurosci. 2013;14(5):311–321. doi: 10.1038/nrn3484. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Guizzetti M, Zhang X, Goeke C, Gavin DP. Glia and neurodevelopment: focus on fetal alcohol spectrum disorders. Front Pediatr. 2014;2:123. doi: 10.3389/fped.2014.00123. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Walz W. Role of glial cells in the regulation of the brain ion microenvironment. Prog Neurobiol. 1989;33(4):309–333. doi: 10.1016/0301-0082(89)90005-1. [DOI] [PubMed] [Google Scholar]
  • 36.Brookes N. In vitro evidence for the role of glutamate in the CNS toxicity of mercury. Toxicology. 1992;76(3):245–256. doi: 10.1016/0300-483x(92)90193-i. [DOI] [PubMed] [Google Scholar]
  • 37.Dave V, Mullaney KJ, Goderie S, Kimelberg HK, Aschner M. Astrocytes as mediators of methylmercury neurotoxicity: effects on D-aspartate and serotonin uptake. Dev Neurosci. 1994;16(3–4):222–231. doi: 10.1159/000112110. [DOI] [PubMed] [Google Scholar]
  • 38.Allen JW, Mutkus LA, Aschner M. Methylmercury-mediated inhibition of 3H-D-aspartate transport in cultured astrocytes is reversed by the antioxidant catalase. Brain Res. 2001;902(1):92–100. doi: 10.1016/s0006-8993(01)02375-7. [DOI] [PubMed] [Google Scholar]
  • 39.Allen JW, Shanker G, Aschner M. Methylmercury inhibits the in vitro uptake of the glutathione precursor, cystine, in astrocytes, but not in neurons. Brain Res. 2001;894(1):131–140. doi: 10.1016/s0006-8993(01)01988-6. [DOI] [PubMed] [Google Scholar]
  • 40.Shanker G, Allen JW, Mutkus LA, Aschner M. The uptake of cysteine in cultured primary astrocytes and neurons. Brain Res. 2001;902(2):156–163. doi: 10.1016/s0006-8993(01)02342-3. [DOI] [PubMed] [Google Scholar]
  • 41.Shanker G, Aschner M. Methylmercury-induced reactive oxygen species formation in neonatal cerebral astrocytic cultures is attenuated by antioxidants. Brain Res Mol Brain Res. 2003;110(1):85–91. doi: 10.1016/s0169-328x(02)00642-3. [DOI] [PubMed] [Google Scholar]
  • 42.Sakamoto M, Miyamoto K, Wu Z, Nakanishi H. Possible involvement of cathepsin B released by microglia in methylmercury-induced cerebellar pathological changes in the adult rat. Neurosci Lett. 2008;442(3):292–296. doi: 10.1016/j.neulet.2008.07.019. [DOI] [PubMed] [Google Scholar]
  • 43.Choi BH, Lapham LW, Amin-Zaki L, Saleem T. Abnormal neuronal migration, deranged cerebral cortical organization, and diffuse white matter astrocytosis of human fetal brain: a major effect of methylmercury poisoning in utero. J Neuropathol Exp Neurol. 1978;37(6):719–733. doi: 10.1097/00005072-197811000-00001. [DOI] [PubMed] [Google Scholar]
  • 44.Choi BH. Methylmercury poisoning of the developing nervous system: I. Pattern of neuronal migration in the cerebral cortex. Neurotoxicology. 1986;7(2):591–600. [PubMed] [Google Scholar]
  • 45.Castoldi AF, Johansson C, Onishchenko N, Coccini T, Roda E, Vahter M, Ceccatelli S, Manzo L. Human developmental neurotoxicity of methylmercury: impact of variables and risk modifiers. Regul Toxicol Pharmacol. 2008;51(2):201–214. doi: 10.1016/j.yrtph.2008.01.016. [DOI] [PubMed] [Google Scholar]
  • 46.Grandjean P, Weihe P, White RF, Debes F, Araki S, Yokoyama K, Murata K, Sorensen N, Dahl R, Jorgensen PJ. Cognitive deficit in 7-year-old children with prenatal exposure to methylmercury. Neurotoxicol Teratol. 1997;19(6):417–428. doi: 10.1016/s0892-0362(97)00097-4. [DOI] [PubMed] [Google Scholar]
  • 47.Harada M. Minamata disease: methylmercury poisoning in Japan caused by environmental pollution. Crit Rev Toxicol. 1995;25(1):1–24. doi: 10.3109/10408449509089885. [DOI] [PubMed] [Google Scholar]
  • 48.Bello SC. Autism and environmental influences: review and commentary. Rev Environ Health. 2007;22(2):139–156. doi: 10.1515/reveh.2007.22.2.139. [DOI] [PubMed] [Google Scholar]
  • 49.Palmer RF, Blanchard S, Wood R. Proximity to point sources of environmental mercury release as a predictor of autism prevalence. Health Place. 2009;15(1):18–24. doi: 10.1016/j.healthplace.2008.02.001. [DOI] [PubMed] [Google Scholar]
  • 50.Desoto MC, Hitlan RT. Blood levels of mercury are related to diagnosis of autism: a reanalysis of an important data set. J Child Neurol. 2007;22(11):1308–1311. doi: 10.1177/0883073807307111. [DOI] [PubMed] [Google Scholar]
  • 51.Ip P, Wong V, Ho M, Lee J, Wong W. Mercury exposure in children with autistic spectrum disorder: case-control study. J Child Neurol. 2004;19(6):431–434. doi: 10.1177/088307380401900606. [DOI] [PubMed] [Google Scholar]
  • 52.Bjorklund O, Kahlstrom J, Salmi P, Ogren SO, Vahter M, Chen JF, Fredholm BB, Dare E. The effects of methylmercury on motor activity are sex- and age-dependent, and modulated by genetic deletion of adenosine receptors and caffeine administration. Toxicology. 2007;241(3):119–133. doi: 10.1016/j.tox.2007.08.092. [DOI] [PubMed] [Google Scholar]
  • 53.Carratu MR, Borracci P, Coluccia A, Giustino A, Renna G, Tomasini MC, Raisi E, Antonelli T, Cuomo V, Mazzoni E, Ferraro L. Acute exposure to methylmercury at two developmental windows: focus on neurobehavioral and neurochemical effects in rat offspring. Neuroscience. 2006;141(3):1619–1629. doi: 10.1016/j.neuroscience.2006.05.017. [DOI] [PubMed] [Google Scholar]
  • 54.Dare E, Fetissov S, Hokfelt T, Hall H, Ogren SO, Ceccatelli S. Effects of prenatal exposure to methylmercury on dopamine-mediated locomotor activity and dopamine D2 receptor binding. Naunyn Schmiedebergs Arch Pharmacol. 2003;367(5):500–508. doi: 10.1007/s00210-003-0716-5. [DOI] [PubMed] [Google Scholar]
  • 55.Sakamoto M, Kakita A, Wakabayashi K, Takahashi H, Nakano A, Akagi H. Evaluation of changes in methylmercury accumulation in the developing rat brain and its effects: a study with consecutive and moderate dose exposure throughout gestation and lactation periods. Brain Res. 2002;949(1–2):51–59. doi: 10.1016/s0006-8993(02)02964-5. [DOI] [PubMed] [Google Scholar]
  • 56.Paletz EM, Craig-Schmidt MC, Newland MC. Gestational exposure to methylmercury and n-3 fatty acids: effects on high- and low-rate operant behavior in adulthood. Neurotoxicol Teratol. 2006;28(1):59–73. doi: 10.1016/j.ntt.2005.11.003. [DOI] [PubMed] [Google Scholar]
  • 57.Radonjic M, Cappaert NL, de Vries EF, de Esch CE, Kuper FC, van Waarde A, Dierckx RA, Wadman WJ, Wolterbeek AP, Stierum RH, de Groot DM. Delay and impairment in brain development and function in rat offspring after maternal exposure to methylmercury. Toxicol Sci. 2013;133(1):112–124. doi: 10.1093/toxsci/kft024. [DOI] [PubMed] [Google Scholar]
  • 58.Rodier PM, Aschner M, Sager PR. Mitotic arrest in the developing CNS after prenatal exposure to methylmercury. Neurobehav Toxicol Teratol. 1984;6(5):379–385. [PubMed] [Google Scholar]
  • 59.Burke K, Cheng Y, Li B, Petrov A, Joshi P, Berman RF, Reuhl KR, DiCicco-Bloom E. Methylmercury elicits rapid inhibition of cell proliferation in the developing brain and decreases cell cycle regulator, cyclin E. Neurotoxicology. 2006;27(6):970–981. doi: 10.1016/j.neuro.2006.09.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Sokolowski K, Falluel-Morel A, Zhou X, DiCicco-Bloom E. Methylmercury (MeHg) elicits mitochondrial-dependent apoptosis in developing hippocampus and acts at low exposures. Neurotoxicology. 2011;32(5):535–544. doi: 10.1016/j.neuro.2011.06.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Engleson G, Herner T. Alkyl mercury poisoning. Acta Paediatr. 1952;41(3):289–294. doi: 10.1111/j.1651-2227.1952.tb17033.x. [DOI] [PubMed] [Google Scholar]
  • 62.Harada M. Congenital Minamata disease: intrauterine methylmercury poisoning. Teratology. 1978;18(2):285–288. doi: 10.1002/tera.1420180216. [DOI] [PubMed] [Google Scholar]
  • 63.Cordier S, Garel M, Mandereau L, Morcel H, Doineau P, Gosme-Seguret S, Josse D, White R, Amiel-Tison C. Neurodevelopmental investigations among methylmercury-exposed children in French Guiana. Environ Res. 2002;89(1):1–11. doi: 10.1006/enrs.2002.4349. [DOI] [PubMed] [Google Scholar]
  • 64.Steuerwald U, Weihe P, Jorgensen PJ, Bjerve K, Brock J, Heinzow B, Budtz-Jorgensen E, Grandjean P. Maternal seafood diet, methylmercury exposure, and neonatal neurologic function. J Pediatr. 2000;136(5):599–605. doi: 10.1067/mpd.2000.102774. [DOI] [PubMed] [Google Scholar]
  • 65.Faustman EM, Ponce RA, Ou YC, Mendoza MA, Lewandowski T, Kavanagh T. Investigations of methylmercury-induced alterations in neurogenesis. Environ Health Perspect. 2002;110(Suppl 5):859–864. doi: 10.1289/ehp.02110s5859. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Choi BH. The effects of methylmercury on the developing brain. Prog Neurobiol. 1989;32(6):447–470. doi: 10.1016/0301-0082(89)90018-x. [DOI] [PubMed] [Google Scholar]
  • 67.Kunimoto M, Suzuki T. Migration of granule neurons in cerebellar organotypic cultures is impaired by methylmercury. Neurosci Lett. 1997;226(3):183–186. doi: 10.1016/s0304-3940(97)00273-5. [DOI] [PubMed] [Google Scholar]
  • 68.Miura K, Himeno S, Koide N, Imura N. Effects of methylmercury and inorganic mercury on the growth of nerve fibers in cultured chick dorsal root ganglia. Tohoku J Exp Med. 2000;192(3):195–210. doi: 10.1620/tjem.192.195. [DOI] [PubMed] [Google Scholar]
  • 69.Nagashima K, Fujii Y, Tsukamoto T, Nukuzuma S, Satoh M, Fujita M, Fujioka Y, Akagi H. Apoptotic process of cerebellar degeneration in experimental methylmercury intoxication of rats. Acta Neuropathol. 1996;91(1):72–77. doi: 10.1007/s004010050394. [DOI] [PubMed] [Google Scholar]
  • 70.Nagashima K. A review of experimental methylmercury toxicity in rats: neuropathology and evidence for apoptosis. Toxicol Pathol. 1997;25(6):624–631. doi: 10.1177/019262339702500613. [DOI] [PubMed] [Google Scholar]
  • 71.Castoldi AF, Barni S, Turin I, Gandini C, Manzo L. Early acute necrosis, delayed apoptosis and cytoskeletal breakdown in cultured cerebellar granule neurons exposed to methylmercury. J Neurosci Res. 2000;59(6):775–787. doi: 10.1002/(SICI)1097-4547(20000315)59:6<775::AID-JNR10>3.0.CO;2-T. [DOI] [PubMed] [Google Scholar]
  • 72.Miura K, Imura N. Mechanism of methylmercury cytotoxicity. Crit Rev Toxicol. 1987;18(3):161–188. doi: 10.3109/10408448709089860. [DOI] [PubMed] [Google Scholar]
  • 73.Ferraro L, Tomasini MC, Tanganelli S, Mazza R, Coluccia A, Carratu MR, Gaetani S, Cuomo V, Antonelli T. Developmental exposure to methylmercury elicits early cell death in the cerebral cortex and long-term memory deficits in the rat. Int J Dev Neurosci. 2009;27(2):165–174. doi: 10.1016/j.ijdevneu.2008.11.004. [DOI] [PubMed] [Google Scholar]
  • 74.Bulleit RF, Cui H. Methylmercury antagonizes the survival-promoting activity of insulin-like growth factor on developing cerebellar granule neurons. Toxicol Appl Pharmacol. 1998;153(2):161–168. doi: 10.1006/taap.1998.8561. [DOI] [PubMed] [Google Scholar]
  • 75.Fonfria E, Dare E, Benelli M, Sunol C, Ceccatelli S. Translocation of apoptosis-inducing factor in cerebellar granule cells exposed to neurotoxic agents inducing oxidative stress. Eur J Neurosci. 2002;16(10):2013–2016. doi: 10.1046/j.1460-9568.2002.02269.x. [DOI] [PubMed] [Google Scholar]
  • 76.Tamm C, Duckworth J, Hermanson O, Ceccatelli S. High susceptibility of neural stem cells to methylmercury toxicity: effects on cell survival and neuronal differentiation. J Neurochem. 2006;97(1):69–78. doi: 10.1111/j.1471-4159.2006.03718.x. [DOI] [PubMed] [Google Scholar]
  • 77.Orrenius S, Nicotera P, Zhivotovsky B. Cell death mechanisms and their implications in toxicology. Toxicol Sci. 2011;119(1):3–19. doi: 10.1093/toxsci/kfq268. [DOI] [PubMed] [Google Scholar]
  • 78.Orrenius S, Zhivotovsky B, Nicotera P. Regulation of cell death: the calcium-apoptosis link. Nat Rev Mol Cell Biol. 2003;4(7):552–565. doi: 10.1038/nrm1150. [DOI] [PubMed] [Google Scholar]
  • 79.Marty MS, Atchison WD. Pathways mediating Ca2+ entry in rat cerebellar granule cells following in vitro exposure to methyl mercury. Toxicol Appl Pharmacol. 1997;147(2):319–330. doi: 10.1006/taap.1997.8262. [DOI] [PubMed] [Google Scholar]
  • 80.Marty MS, Atchison WD. Elevations of intracellular Ca2+ as a probable contributor to decreased viability in cerebellar granule cells following acute exposure to methylmercury. Toxicol Appl Pharmacol. 1998;150(1):98–105. doi: 10.1006/taap.1998.8383. [DOI] [PubMed] [Google Scholar]
  • 81.Sirois JE, Atchison WD. Methylmercury affects multiple subtypes of calcium channels in rat cerebellar granule cells. Toxicol Appl Pharmacol. 2000;167(1):1–11. doi: 10.1006/taap.2000.8967. [DOI] [PubMed] [Google Scholar]
  • 82.Sakamoto M, Ikegami N, Nakano A. Protective effects of Ca2+ channel blockers against methyl mercury toxicity. Pharmacol Toxicol. 1996;78(3):193–199. doi: 10.1111/j.1600-0773.1996.tb00203.x. [DOI] [PubMed] [Google Scholar]
  • 83.Sarafian TA. Methyl mercury increases intracellular Ca2+ and inositol phosphate levels in cultured cerebellar granule neurons. J Neurochem. 1993;61(2):648–657. doi: 10.1111/j.1471-4159.1993.tb02169.x. [DOI] [PubMed] [Google Scholar]
  • 84.Gasso S, Cristofol RM, Selema G, Rosa R, Rodriguez-Farre E, Sanfeliu C. Antioxidant compounds and Ca(2+) pathway blockers differentially protect against methylmercury and mercuric chloride neurotoxicity. J Neurosci Res. 2001;66(1):135–145. doi: 10.1002/jnr.1205. [DOI] [PubMed] [Google Scholar]
  • 85.Okazaki E, Oyama Y, Chikahisa L, Nagano T, Katayama N, Sakamoto M. Fluorescent estimation on cytotoxicity of methylmercury in dissociated rat cerebellar neurons: its comparison with ionomycin. Environ Toxicol Pharmacol. 1997;3(4):237–244. doi: 10.1016/s1382-6689(97)00017-3. [DOI] [PubMed] [Google Scholar]
  • 86.Oyama Y, Tomiyoshi F, Ueno S, Furukawa K, Chikahisa L. Methylmercury-induced augmentation of oxidative metabolism in cerebellar neurons dissociated from the rats: its dependence on intracellular Ca2+ Brain Res. 1994;660(1):154–157. doi: 10.1016/0006-8993(94)90849-4. [DOI] [PubMed] [Google Scholar]
  • 87.Shanker G, Aschner M. Identification and characterization of uptake systems for cystine and cysteine in cultured astrocytes and neurons: evidence for methylmercury-targeted disruption of astrocyte transport. J Neurosci Res. 2001;66(5):998–1002. doi: 10.1002/jnr.10066. [DOI] [PubMed] [Google Scholar]
  • 88.Manfroi CB, Schwalm FD, Cereser V, Abreu F, Oliveira A, Bizarro L, Rocha JB, Frizzo ME, Souza DO, Farina M. Maternal milk as methylmercury source for suckling mice: neurotoxic effects involved with the cerebellar glutamatergic system. Toxicol Sci. 2004;81(1):172–178. doi: 10.1093/toxsci/kfh201. [DOI] [PubMed] [Google Scholar]
  • 89.Miyamoto K, Nakanishi H, Moriguchi S, Fukuyama N, Eto K, Wakamiya J, Murao K, Arimura K, Osame M. Involvement of enhanced sensitivity of N-methyl-D-aspartate receptors in vulnerability of developing cortical neurons to methylmercury neurotoxicity. Brain Res. 2001;901(1–2):252–258. doi: 10.1016/s0006-8993(01)02281-8. [DOI] [PubMed] [Google Scholar]
  • 90.Reynolds IJ, Hastings TG. Glutamate induces the production of reactive oxygen species in cultured forebrain neurons following NMDA receptor activation. J Neurosci. 1995;15(5 Pt 1):3318–3327. doi: 10.1523/JNEUROSCI.15-05-03318.1995. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 91.Stringari J, Nunes AK, Franco JL, Bohrer D, Garcia SC, Dafre AL, Milatovic D, Souza DO, Rocha JB, Aschner M, Farina M. Prenatal methylmercury exposure hampers glutathione antioxidant system ontogenesis and causes long-lasting oxidative stress in the mouse brain. Toxicol Appl Pharmacol. 2008;227(1):147–154. doi: 10.1016/j.taap.2007.10.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 92.Franco JL, Teixeira A, Meotti FC, Ribas CM, Stringari J, Garcia Pomblum SC, Moro AM, Bohrer D, Bairros AV, Dafre AL, Santos AR, Farina M. Cerebellar thiol status and motor deficit after lactational exposure to methylmercury. Environ Res. 2006;102(1):22–28. doi: 10.1016/j.envres.2006.02.003. [DOI] [PubMed] [Google Scholar]
  • 93.Fahrion JK, Komuro Y, Li Y, Ohno N, Littner Y, Raoult E, Galas L, Vaudry D, Komuro H. Rescue of neuronal migration deficits in a mouse model of fetal Minamata disease by increasing neuronal Ca2+ spike frequency. Proc Natl Acad Sci U S A. 2012;109(13):5057–5062. doi: 10.1073/pnas.1120747109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94.Sass JB, Haselow DT, Silbergeld EK. Methylmercury-induced decrement in neuronal migration may involve cytokine-dependent mechanisms: a novel method to assess neuronal movement in vitro. Toxicol Sci. 2001;63(1):74–81. doi: 10.1093/toxsci/63.1.74. [DOI] [PubMed] [Google Scholar]
  • 95.Mancini JD, Autio DM, Atchison WD. Continuous exposure to low concentrations of methylmercury impairs cerebellar granule cell migration in organotypic slice culture. Neurotoxicology. 2009;30(2):203–208. doi: 10.1016/j.neuro.2008.12.010. [DOI] [PubMed] [Google Scholar]
  • 96.Limke TL, Atchison WD. Acute exposure to methylmercury opens the mitochondrial permeability transition pore in rat cerebellar granule cells. Toxicol Appl Pharmacol. 2002;178(1):52–61. doi: 10.1006/taap.2001.9327. [DOI] [PubMed] [Google Scholar]
  • 97.Hare MF, Atchison WD. Methylmercury mobilizes Ca++ from intracellular stores sensitive to inositol 1,4,5-trisphosphate in NG108–15 cells. J Pharmacol Exp Ther. 1995;272(3):1016–1023. [PubMed] [Google Scholar]
  • 98.Hare MF, McGinnis KM, Atchison WD. Methylmercury increases intracellular concentrations of Ca++ and heavy metals in NG108–15 cells. J Pharmacol Exp Ther. 1993;266(3):1626–1635. [PubMed] [Google Scholar]
  • 99.Farina M, Campos F, Vendrell I, Berenguer J, Barzi M, Pons S, Sunol C. Probucol increases glutathione peroxidase-1 activity and displays long-lasting protection against methylmercury toxicity in cerebellar granule cells. Toxicol Sci. 2009;112(2):416–426. doi: 10.1093/toxsci/kfp219. [DOI] [PubMed] [Google Scholar]
  • 100.Liddell JR, Dringen R, Crack PJ, Robinson SR. Glutathione peroxidase 1 and a high cellular glutathione concentration are essential for effective organic hydroperoxide detoxification in astrocytes. Glia. 2006;54(8):873–879. doi: 10.1002/glia.20433. [DOI] [PubMed] [Google Scholar]
  • 101.Komuro H, Rakic P. Intracellular Ca2+ fluctuations modulate the rate of neuronal migration. Neuron. 1996;17(2):275–285. doi: 10.1016/s0896-6273(00)80159-2. [DOI] [PubMed] [Google Scholar]
  • 102.Kumada T, Komuro H. Completion of neuronal migration regulated by loss of Ca(2+) transients. Proc Natl Acad Sci U S A. 2004;101(22):8479–8484. doi: 10.1073/pnas.0401000101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 103.Jaconi ME, Theler JM, Schlegel W, Appel RD, Wright SD, Lew PD. Multiple elevations of cytosolic-free Ca2+ in human neutrophils: initiation by adherence receptors of the integrin family. J Cell Biol. 1991;112(6):1249–1257. doi: 10.1083/jcb.112.6.1249. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 104.Marks PW, Maxfield FR. Transient increases in cytosolic free calcium appear to be required for the migration of adherent human neutrophils. J Cell Biol. 1990;110(1):43–52. doi: 10.1083/jcb.110.1.43. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 105.Hofmann F, Biel M, Flockerzi V. Molecular basis for Ca2+ channel diversity. Annu Rev Neurosci. 1994;17:399–418. doi: 10.1146/annurev.ne.17.030194.002151. [DOI] [PubMed] [Google Scholar]
  • 106.Reuter H. Diversity and function of presynaptic calcium channels in the brain. Curr Opin Neurobiol. 1996;6(3):331–337. doi: 10.1016/s0959-4388(96)80116-4. [DOI] [PubMed] [Google Scholar]
  • 107.Randall A, Tsien RW. Pharmacological dissection of multiple types of Ca2+ channel currents in rat cerebellar granule neurons. J Neurosci. 1995;15(4):2995–3012. doi: 10.1523/JNEUROSCI.15-04-02995.1995. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 108.Komuro H, Rakic P. Selective role of N-type calcium channels in neuronal migration. Science. 1992;257(5071):806–809. doi: 10.1126/science.1323145. [DOI] [PubMed] [Google Scholar]
  • 109.Komuro H, Rakic P. Orchestration of neuronal migration by activity of ion channels, neurotransmitter receptors, and intracellular Ca2+ fluctuations. J Neurobiol. 1998;37(1):110–130. [PubMed] [Google Scholar]
  • 110.Shafer TJ. Effects of Cd2+, Pb2+ and CH3Hg+ on high voltage-activated calcium currents in pheochromocytoma (PC12) cells: potency, reversibility, interactions with extracellular Ca2+ and mechanisms of block. Toxicol Lett. 1998;99(3):207–221. doi: 10.1016/s0378-4274(98)00225-2. [DOI] [PubMed] [Google Scholar]
  • 111.Shafer TJ, Atchison WD. Methylmercury blocks N- and L-type Ca++ channels in nerve growth factor-differentiated pheochromocytoma (PC12) cells. J Pharmacol Exp Ther. 1991;258(1):149–157. [PubMed] [Google Scholar]
  • 112.Shafer TJ, Meacham CA, Barone S., Jr Effects of prolonged exposure to nanomolar concentrations of methylmercury on voltage-sensitive sodium and calcium currents in PC12 cells. Brain Res Dev Brain Res. 2002;136(2):151–164. doi: 10.1016/s0165-3806(02)00360-7. [DOI] [PubMed] [Google Scholar]
  • 113.Bradford AB, Mancini JD, Atchison WD. Methylmercury-Dependent Increases in Fluo4 Fluorescence in Neonatal Rat Cerebellar Slices Depend on Granule Cell Migrational Stage and GABAA Receptor Modulation. J Pharmacol Exp Ther. 2016;356(1):2–12. doi: 10.1124/jpet.115.226761. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 114.Yuan Y, Atchison WD. Methylmercury acts at multiple sites to block hippocampal synaptic transmission. J Pharmacol Exp Ther. 1995;275(3):1308–1316. [PubMed] [Google Scholar]
  • 115.Yuan Y, Atchison WD. Action of methylmercury on GABA(A) receptor-mediated inhibitory synaptic transmission is primarily responsible for its early stimulatory effects on hippocampal CA1 excitatory synaptic transmission. J Pharmacol Exp Ther. 1997;282(1):64–73. [PubMed] [Google Scholar]
  • 116.Yuan Y, Atchison WD. Methylmercury differentially affects GABA(A) receptor-mediated spontaneous IPSCs in Purkinje and granule cells of rat cerebellar slices. J Physiol. 2003;550(Pt 1):191–204. doi: 10.1113/jphysiol.2003.040543. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 117.Hogberg HT, Kinsner-Ovaskainen A, Coecke S, Hartung T, Bal-Price AK. mRNA expression is a relevant tool to identify developmental neurotoxicants using an in vitro approach. Toxicol Sci. 2010;113(1):95–115. doi: 10.1093/toxsci/kfp175. [DOI] [PubMed] [Google Scholar]
  • 118.Zimmermann KC, Bonzon C, Green DR. The machinery of programmed cell death. Pharmacol Ther. 2001;92(1):57–70. doi: 10.1016/s0163-7258(01)00159-0. [DOI] [PubMed] [Google Scholar]
  • 119.Fujimura M, Usuki F. Methylmercury causes neuronal cell death through the suppression of the TrkA pathway: in vitro and in vivo effects of TrkA pathway activators. Toxicol Appl Pharmacol. 2015;282(3):259–266. doi: 10.1016/j.taap.2014.12.008. [DOI] [PubMed] [Google Scholar]
  • 120.Kuida K, Zheng TS, Na S, Kuan C, Yang D, Karasuyama H, Rakic P, Flavell RA. Decreased apoptosis in the brain and premature lethality in CPP32-deficient mice. Nature. 1996;384(6607):368–372. doi: 10.1038/384368a0. [DOI] [PubMed] [Google Scholar]
  • 121.Fernando P, Brunette S, Megeney LA. Neural stem cell differentiation is dependent upon endogenous caspase 3 activity. FASEB J. 2005;19(12):1671–1673. doi: 10.1096/fj.04-2981fje. [DOI] [PubMed] [Google Scholar]
  • 122.Deng W, Aimone JB, Gage FH. New neurons and new memories: how does adult hippocampal neurogenesis affect learning and memory? Nat Rev Neurosci. 2010;11(5):339–350. doi: 10.1038/nrn2822. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 123.Sokolowski K, Obiorah M, Robinson K, McCandlish E, Buckley B, DiCicco-Bloom E. Neural stem cell apoptosis after low-methylmercury exposures in postnatal hippocampus produce persistent cell loss and adolescent memory deficits. Dev Neurobiol. 2013;73(12):936–949. doi: 10.1002/dneu.22119. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 124.Dare E, Gotz ME, Zhivotovsky B, Manzo L, Ceccatelli S. Antioxidants J811 and 17beta-estradiol protect cerebellar granule cells from methylmercury-induced apoptotic cell death. J Neurosci Res. 2000;62(4):557–565. doi: 10.1002/1097-4547(20001115)62:4<557::AID-JNR10>3.0.CO;2-9. [DOI] [PubMed] [Google Scholar]
  • 125.Sakaue M, Okazaki M, Hara S. Very low levels of methylmercury induce cell death of cultured rat cerebellar neurons via calpain activation. Toxicology. 2005;213(1–2):97–106. doi: 10.1016/j.tox.2005.05.013. [DOI] [PubMed] [Google Scholar]
  • 126.Tofighi R, Johansson C, Goldoni M, Ibrahim WN, Gogvadze V, Mutti A, Ceccatelli S. Hippocampal neurons exposed to the environmental contaminants methylmercury and polychlorinated biphenyls undergo cell death via parallel activation of calpains and lysosomal proteases. Neurotox Res. 2011;19(1):183–194. doi: 10.1007/s12640-010-9159-1. [DOI] [PubMed] [Google Scholar]
  • 127.Wang KK. Calpain and caspase: can you tell the difference? Trends Neurosci. 2000;23(1):20–26. doi: 10.1016/s0166-2236(99)01479-4. [DOI] [PubMed] [Google Scholar]
  • 128.Camins A, Verdaguer E, Folch J, Pallas M. Involvement of calpain activation in neurodegenerative processes. CNS Drug Rev. 2006;12(2):135–148. doi: 10.1111/j.1527-3458.2006.00135.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 129.Johnson GV, Guttmann RP. Calpains: intact and active? Bioessays. 1997;19(11):1011–1018. doi: 10.1002/bies.950191111. [DOI] [PubMed] [Google Scholar]
  • 130.Raheja G, Gill KD. Calcium homeostasis and dichlorvos induced neurotoxicity in rat brain. Mol Cell Biochem. 2002;232(1–2):13–18. doi: 10.1023/a:1014873031013. [DOI] [PubMed] [Google Scholar]
  • 131.Chan SL, Mattson MP. Caspase and calpain substrates: roles in synaptic plasticity and cell death. J Neurosci Res. 1999;58(1):167–190. [PubMed] [Google Scholar]
  • 132.Sorimachi H, Ishiura S, Suzuki K. Structure and physiological function of calpains. Biochem J. 1997;328(Pt 3):721–732. doi: 10.1042/bj3280721. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 133.Poulain FE, Sobel A. The microtubule network and neuronal morphogenesis: Dynamic and coordinated orchestration through multiple players. Mol Cell Neurosci. 2010;43(1):15–32. doi: 10.1016/j.mcn.2009.07.012. [DOI] [PubMed] [Google Scholar]
  • 134.Feng J. Microtubule: a common target for parkin and Parkinson’s disease toxins. Neuroscientist. 2006;12(6):469–476. doi: 10.1177/1073858406293853. [DOI] [PubMed] [Google Scholar]
  • 135.Sager PR, Doherty RA, Olmsted JB. Interaction of methylmercury with microtubules in cultured cells and in vitro. Exp Cell Res. 1983;146(1):127–137. doi: 10.1016/0014-4827(83)90331-2. [DOI] [PubMed] [Google Scholar]
  • 136.Vogel DG, Margolis RL, Mottet NK. The effects of methyl mercury binding to microtubules. Toxicol Appl Pharmacol. 1985;80(3):473–486. doi: 10.1016/0041-008x(85)90392-8. [DOI] [PubMed] [Google Scholar]
  • 137.Vogel DG, Margolis RL, Mottet NK. Analysis of methyl mercury binding sites on tubulin subunits and microtubules. Pharmacol Toxicol. 1989;64(2):196–201. doi: 10.1111/j.1600-0773.1989.tb00630.x. [DOI] [PubMed] [Google Scholar]
  • 138.Vendrell I, Carrascal M, Campos F, Abian J, Sunol C. Methylmercury disrupts the balance between phosphorylated and non-phosphorylated cofilin in primary cultures of mice cerebellar granule cells. A proteomic study. Toxicol Appl Pharmacol. 2010;242(1):109–118. doi: 10.1016/j.taap.2009.09.022. [DOI] [PubMed] [Google Scholar]
  • 139.Bravo-Cordero JJ, Magalhaes MA, Eddy RJ, Hodgson L, Condeelis J. Functions of cofilin in cell locomotion and invasion. Nat Rev Mol Cell Biol. 2013;14(7):405–415. doi: 10.1038/nrm3609. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 140.Xu M, Yan C, Tian Y, Yuan X, Shen X. Effects of low level of methylmercury on proliferation of cortical progenitor cells. Brain Res. 2010;1359:272–280. doi: 10.1016/j.brainres.2010.08.069. [DOI] [PubMed] [Google Scholar]
  • 141.Stankiewicz TR, Linseman DA. Rho family GTPases: key players in neuronal development, neuronal survival, and neurodegeneration. Front Cell Neurosci. 2014;8:314. doi: 10.3389/fncel.2014.00314. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 142.Jaffe AB, Hall A. Rho GTPases: biochemistry and biology. Annu Rev Cell Dev Biol. 2005;21:247–269. doi: 10.1146/annurev.cellbio.21.020604.150721. [DOI] [PubMed] [Google Scholar]
  • 143.Fujimura M, Usuki F, Sawada M, Rostene W, Godefroy D, Takashima A. Methylmercury exposure downregulates the expression of Racl and leads to neuritic degeneration and ultimately apoptosis in cerebrocortical neurons. Neurotoxicology. 2009;30(1):16–22. doi: 10.1016/j.neuro.2008.10.002. [DOI] [PubMed] [Google Scholar]
  • 144.Tan HB, Zhong YS, Cheng Y, Shen X. Rho/ROCK pathway and neural regeneration: a potential therapeutic target for central nervous system and optic nerve damage. Int J Ophthalmol. 2011;4(6):652–657. doi: 10.3980/j.issn.2222-3959.2011.06.16. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 145.Fujimura M, Usuki F, Kawamura M, Izumo S. Inhibition of the Rho/ROCK pathway prevents neuronal degeneration in vitro and in vivo following methylmercury exposure. Toxicol Appl Pharmacol. 2011;250(1):1–9. doi: 10.1016/j.taap.2010.09.011. [DOI] [PubMed] [Google Scholar]
  • 146.Ahmed S, Gan HT, Lam CS, Poonepalli A, Ramasamy S, Tay Y, Tham M, Yu YH. Transcription factors and neural stem cell self-renewal, growth and differentiation. Cell Adh Migr. 2009;3(4):412–424. doi: 10.4161/cam.3.4.8803. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 147.Tamm C, Duckworth JK, Hermanson O, Ceccatelli S. Methylmercury inhibits differentiation of rat neural stem cells via Notch signalling. Neuroreport. 2008;19(3):339–343. doi: 10.1097/WNR.0b013e3282f50ca4. [DOI] [PubMed] [Google Scholar]
  • 148.Levy OA, Lah JJ, Levey AI. Notch signaling inhibits PC12 cell neurite outgrowth via RBP-J-dependent and -independent mechanisms. Dev Neurosci. 2002;24(1):79–88. doi: 10.1159/000064948. [DOI] [PubMed] [Google Scholar]
  • 149.Bose R, Onishchenko N, Edoff K, Janson Lang AM, Ceccatelli S. Inherited effects of low-dose exposure to methylmercury in neural stem cells. Toxicol Sci. 2012;130(2):383–390. doi: 10.1093/toxsci/kfs257. [DOI] [PubMed] [Google Scholar]
  • 150.Theunissen PT, Pennings JL, Robinson JF, Claessen SM, Kleinjans JC, Piersma AH. Time-response evaluation by transcriptomics of methylmercury effects on neural differentiation of murine embryonic stem cells. Toxicol Sci. 2011;122(2):437–447. doi: 10.1093/toxsci/kfr134. [DOI] [PubMed] [Google Scholar]
  • 151.Fujimura M, Usuki F. Low concentrations of methylmercury inhibit neural progenitor cell proliferation associated with up-regulation of glycogen synthase kinase 3beta and subsequent degradation of cyclin E in rats. Toxicol Appl Pharmacol. 2015;288(1):19–25. doi: 10.1016/j.taap.2015.07.006. [DOI] [PubMed] [Google Scholar]

RESOURCES