1. Introduction
The nervous system is the major controlling, regulatory and communicating system in the body. In vertebrates, it consists of a central nervous system (CNS, brain) and a peripheral nervous system (spinal cord and ganglia). The CNS controls voluntary movements such as walking, speaking, as well as involuntary movements as breathing and reflex actions. It also controls learning, emotions and metabolism. Therefore, any type of damage to the nervous system (mechanical trauma or poisoning) will impact biological functions that might be essential for life, such as breathing or heart pumping (Tortora and Derrickson, 2018).
Accordingly, the nervous system is well preserved. Protective membranes known as the meninges surround the brain and spinal cord, and both float in a crystal-clear cerebrospinal fluid. The central nervous system lies largely within the axial skeleton, wherein the brain is encased in a bony vault, the neurocranium, while the cylindrical and elongated spinal cord lies in the vertebral canal, which is formed by successive vertebrae connected by dense ligaments (Tortora and Derrickson, 2018). In addition, endothelial cells in the majority of the CNS (excluding the circumventricular organs) for a specialized barrier, called the blood-brain barrier, which prevents the crossing of solutes from the blood into the extracellular fluid of the nervous system (Gupta et al., 2019). Some molecules cross by passive diffusion, especially small and non-polar ones. There are transporters that allow the entry of certain molecules, such as glucose and ions (Banks, 2016).
Despite the existence of the BBB, many toxicants reach the nervous system, enter the neurons and glial cells and trigger molecular alterations such as oxidative stress, mitochondrial damage, apoptosis, inhibition or stimulation of signaling pathways, DNA damage and protein and lipid oxidation. These effects can cause abnormal neuronal function or even lead to cell death. Among these toxicants are select metals, which have been described for several years as potentially neurotoxic (Caito and Aschner, 2015). They can enter the brain through transporters such as the divalent metal transporter DMT and zinc transporter ZIP4 or secondary to microvascular damage and opening of interendothelial tight junctions (Zheng et al., 2003). We are constantly exposed to essential and non-essential metals through water, food, air, pharmaceutics, cosmetics and some subjects are exposed in their work environment. In addition, accidents such as the disruption of mining tailings dams as in Mariana and Brumadinho (Brazil) cause a tragic ecological impact, besides leading to diseases to the exposed populations. Non-essential metals possess higher neurotoxicological risk than essential metals, therefore dose and time of exposure are important factors to be considered. Although the great number of studies on the metals neurotoxicology, not all mechanisms have been fully deciphered.
More recently, with the advance of nanotechnology, metallic nanoparticles (NPs) gained attention because of their unique characteristics and applications in industry. The increasing NPs applications lead inevitably to their release into the environment, contaminating water, soil, air and food (Peralta-Videa et al., 2011). In addition, we are constantly in contact with products that contain NPs, such as deodorants and clothes with Ag-NPs in their composition. Notably, an increasing number of studies have evidenced that metallic NPs compose a neurotoxic hazard and understanding the impact of size, charge, shape, ion release and coating on the biological outcomes is urgent (Teleanu et al., 2019).
In order to increase the understanding on metal- and NPs-induced neurotoxicology, alternative and complementary animal models have been used. Invertebrates as the nematode Caenorhabditis elegans have been successfully applied to study basic biological processes such as aging, oxidative stress, and nervous system function (Queirós et al., 2019; Wu et al., 2019). C. elegans has a small size, simple structure, short lifespan, accessibility to genetic manipulation, and the conserved biological pathways in relation to mammals. It allows an ecotoxicological approach because of its presence in the environment, and largely the findings can be extrapolated to humans because of the great genetic homology between the nematode and mammals (Avila et al., 2012).
Indeed, this organism has a simpler nervous system, but it includes almost all the neurotransmitters systems present in humans. The worm has been studied to such an extent that a neuronal wiring diagram is available. The nervous system has high sensitivity to metals and metallic NPs and allows for a wide range of assays, from physiological to behavioral, biochemical and molecular (such as transcriptomics and proteomics) in a faster and more bioethical and economical manner (Avila et al., 2012; Queirós et al., 2019; Sedensky and Morgan, 2018; Soares et al., 2017). The use of mutants and transgenic supports the elucidation of mechanisms that were unclear or inconceivable to investigate in other models. In this context, we have structured the present review to shed light on applications and to update on neurotoxic mechanisms uncovered by studies in this animal model. Our focus is centered on several metals and metallic NPs that cause neurotoxicity in C. elegans, but it is noteworthy that many toxicants have yet to be tested yet in this model.
2. C. elegans neuronal system
The C. elegans nervous system is relatively simple and the whole developmental cell lineage, the anatomical location, structure, and network connections of each neuron have been characterized (Sulston, 1983). A typical adult organism has 302 neurons in adult hermaphrodites (383 in males), divided into 118 morphologically distinct classes and 56 glial cells. Approximately 7000 synapses that compose 2 independent nervous systems have been described, in addition to a large somatic system with 282 neurons and a small pharyngeal system with 20 neurons (White et al., 1986). C. elegans is the only organism that has a fully characterized neuronal wiring diagram where all neurons and the major interactions between them are well characterized and mapped.
The biochemistry of the C. elegans nervous system is highly conserved with mammals—as the worms express similar ion channels, receptors, vesicular transporters, and other synaptic components (Bargmann, 1998). Neurotransmitter synthesis, neuronal vesicles, neurotransmitter release, receptor binding and inactivation (by reuptake or degradation), are well understood, which is important for the study of possible damage to the nervous system.
Nematodes contain a variety of classic neurotransmitters such as dopamine (DA), aminobutyric acid (GABA), glutamate (GLU), acetylcholine, serotonin (5-hydroxytryptamine), and neuropeptides. Proteins related to the nervous system can be co-expressed with green fluorescent protein (GFP) and therefore neuronal morphology can be visualized in vivo (Chalfie et al., 1994). Neuronal excitability can be followed by live calcium imaging (Nguyen et al., 2016), which allows for associations between neuronal activity and behavioral phenotypes. Moreover, its ease of genetic manipulations allows the identification of genes significant for neuronal formation, migration, activity, or other functions (Brenner, 1973). On the other hand, worms lack epinephrine, norepinephrine and histamine signaling, and some significant differences have been described in sodium-dependent channels (Bargmann, 1998; Goodman et al., 1998). In addition, nematodes lack a blood-brain barrier, and once the molecules are absorbed, they can quickly disperse and diffuse into the nervous system (Li et al., 2018; Liberati et al., 2004; Zhao et al., 2016). The structural and functional similarities with the human nervous system, along with other advantages, strengthen the utility of worms as a model in neuroscience and in neurotoxicology. Notably, environmental and occupational exposure to metallic neurotoxicants is a matter of growing concern, since it may have very significant consequences for human health, from impairing neurodevelopment in children to accelerating the evolution of neurodegenerative diseases. The evaluation of the risks associated with the release of metals and metallic NPs in the environment, their modifications (by microorganisms, light, etc.) and bioaccumulation in animals through the use of relevant biological models to assess neurotoxicity are important research objectives (Ortega and Carmona, 2022). Because of the extensive knowledge on C. elegans nervous system and the advantages afforded by mutants, the nematode offers unique perspectives as a neurotoxicological model, which use has been growing recently in the toxicological community. Studies that examine metal- and NPs-induced neurotoxicity are addressed by a wide range of parameters, including behavioral, structural, signaling, and molecular levels. Behavior, even in simple metazoans, depends upon integrated processes at the subcellular, cellular, and organismal level, and thus is susceptible to disruption by a broad spectrum of chemicals. For instance, neuronal systems function can be assessed by simple assays such as locomotion (frequently evaluated by head thrashing and body bends), pharyngeal pumping, defecation, egg laying, chemotaxis and response to a mechanical touch. Well characterized neural circuits control these behaviors and neuronal dysfunction could be a plausible explanation for abnormal outcomes in otherwise anatomically and functionally intact worms. For example, damage of neurons, which innervate muscles of a fully developed vulva (reproductive organ of adult hermaphrodite worms), may result in perturbation of egg laying, reduction of progeny and cause bag of worms (BOW) phenomenon (hatching of the larvae inside parental body) (Pluskota et al., 2009; Scharf et al., 2016). The morphology of the damaged neurons can be visualized by green fluorescent protein (GFP) tagging and identify putative mechanisms, PCR analysis can be done to detect up and down regulated genes. Further investigation of alterations in these endpoints with tools like molecular techniques, pharmacological rescue and transgenic strains can shed light to pathways leading to neurotoxicity and how to attenuate the damages caused by metals and metallic NPs. In this chapter, we will focus on the findings in C. elegans studies focusing on advancing neurotoxicology research of these important toxicants.
3. Metals neurotoxicology in C. elegans
Metallic elements are biologically classified according to their essentiality for human health, and therefore they can be considered essential, inert, or toxic (Crisponi et al., 2012a). To be considered essential, a metal must be present in several tissues, must have important physiological functions and cause irreversible and serious damage to the body when absent (Zoroddu et al., 2019). It should be noted that the essentiality varies according to the species, such as vanadium, which has no known biological effect on humans but is considered essential for algae and fungi (Maret, 2016).
In general, most authors consider that there are currently 10 essential metals: sodium (Na), potassium (K), magnesium (Mg), calcium (Ca), manganese (Mn), iron (Fe), cobalt (Co), copper (Cu), zinc (Zn) and molybdenum (Mo) ( Jomova et al., 2022; Zoroddu et al., 2019). There are several biological functions described for each of these essential metals. Among them we can mention the participation in biochemical reactions (as co-factor or present in prosthetic group), in maintaining electrical charges and osmotic pressure, in mitochondrial respiration, controlling in the production of ROS, maintenance of pairing, stacking and the stability of nucleotide bases and regulation of hormone levels, just to name a few ( Jomova et al., 2022).
Although needed for biological functions, studies point out that excessive intake or exposure to these essential metals can cause toxic effects, including neuronal damage. They can trigger immune response, induce the production of reactive oxygen species (ROS) (such as Fe and Cu by Fenton reaction), alter cell structure and cause cell death. Therefore, it has been established safe levels for intake in order to avoid unwanted effects (Crisponi et al., 2012b).
Toxic metals are defined by their bioaccumulative potential and high toxicity to living organisms, and are also classified as heavy metals, including cadmium (Cd), lead (Pb), nickel (Ni), chromium (Cr), mercury (Hg) and metalloids, such as arsenic (As) (Witkowska et al., 2021). These metals can be taken up by organisms by same transporters for essential metals since they share similar chemical characteristics. They can cause damage to different systems at lower levels than essential metals, including the nervous system. The effects are mediated by diverse mechanisms, causing an increase in ROS, oxidative stress, cellular degeneration, among others (Balali-Mood et al., 2021; Vellingiri et al., 2022; Witkowska et al., 2021).
Metals occur naturally throughout the Earth’s crust, the natural soil erosion allows food sources to be rich in nutrients, including essential metals. Thus, through food, it is possible to acquire adequate concentrations of these nutrients and the organism is tolerable with the small concentrations that are ingested of non-essential metals (Zoroddu et al., 2019). However, anthropogenic activities such as mining and smelting, in addition to industrial production and use, domestic and agricultural use of metals and metal-containing compounds contribute to environmental contamination and human exposure, mainly from heavy metals. The increase in the exposure of these metals also occur due to the use of these elements in medicines and cosmetic products (Balali-Mood et al., 2021; Tchounwou et al., 2012).
3.1. Manganese
Manganese (Mn) is a heavy metal of great abundance in the world. Living beings are readily exposed to this metal due to natural erosion that allows the release of the metal into the air, soil, and water sources. Due to this bioavailability, a wide variety of foods have Mn in their composition, such as leafy vegetables, nuts, brown rice, chickpeas, sweet potatoes, pine nuts, among others (Chen et al., 2015b; USDA, 2019). The wide variety of foods rich in Mn makes food intake the main source of exposure to the metal in humans. However, contamination in drinking water, ingestion of food supplements and exposure to metallurgical industries, polluted air and agricultural inputs are also cited as sources of environmental and occupational exposures and can cause toxicity (Balachandran et al., 2020).
Mn toxicity is of great concern due to the substantial risk of exposure potential damage. It is believed that the neurotoxicity of Mn is caused by the metal’s high capacity to accumulate in astrocytes due to its affinity for the enzyme glutamine synthetase (GS) which is located in these cells (Sidoryk-Wegrzynowicz and Aschner, 2013). Mn exposure to exceedingly high Mn levels may cause a pathological condition described as manganism, which is characterized by dystonia, bradykinesia, and gait disturbance due to irreversible damage to the basal ganglia of the brain (Bowler et al., 2015). Manganism shares molecular mechanisms and symptoms with Parkinson’s Disease (PD) because both cause loss of dopaminergic neurons, impairing dopaminergic function (Gubert et al., 2018). One of the hypotheses is that Mn can cause the oxidation of dopamine, leading to the formation of ROS, in turn, triggering neurodegeneration (Chen et al., 2015c).
In order to further elucidate the Mn neurotoxicological mechanisms, a variety of studies using C. elegans have been done. Using molecular biology tools and genetically modified animals, the studies focused on clarifying the pathways involved in the toxicity of the metal in the neurons. Mn intoxication in C. elegans shares similar effects with those inherent to mammals, including dopaminergic neurodegeneration and oxidative stress. Using this model, it was established that Mn interacts with extracellular dopamine to cause dopaminergic neurodegeneration and that the NADPH dual-oxidase BLI-3 promotes the conversion of extracellular DA into toxic reactive species that are taken up by neurons, leading to neuronal death (Benedetto et al., 2010). There are also genetic models for PD available in C. elegans that present a deletion of DJ-1 related (djr) genes, parkin (pdr-1) and PINK (pink-1) as well, besides mutants expressing the mutated form of alfa-synuclein. All these mutants showed higher sensitivity to Mn, with inherent dopaminergic neurodegeneration (Bornhorst et al., 2014; Chakraborty et al., 2015; Chen et al., 2015b; Gitler et al., 2009).
A recent study used C. elegans as a model to demonstrate the role of Mn in neurodevelopment. Worms were exposed to the metal in the first larval stage (L1) and demonstrated significant deficits in olfactory adaptive learning and memory in first-day adults, as observed in mammals. On the other hand, repeated exposure to Mn in adulthood caused resistance to dopaminergic neurodegeneration (Raj et al., 2021). Another study using the model has proven that the ability of Mn to impair the dopaminergic pathway influences the accumulation of lipids. Worms exposed to Mn for 4 h presented reduced reproduction and had an increase in fat accumulation, a reduction in lipid metabolism and a reduction in dopamine content. The studies also indicated the involvement of dopamine receptors in Mn-induced triacylglycerides accumulation, indicating that Mn interferes with reproduction by blocking lipids transport to eggs, potentiating fat accumulation as a secondary effect following dopaminergic signaling impairment (Gubert et al., 2016, 2018).
New insights into kinetics and therapeutics related to Mn toxicity have also been investigated in the C. elegans model. The molecular similarity in metal absorption and excretion pathways has allowed therapeutic strategies to be evaluated. C. elegans conserves the smf-1, −2 and −3 genes, orthologs to the divalent metal transporters (DMT) in mammals, in addition to expressing the exporter ferroportin (fpn-1.1), which allows an overview of the metabolic state of metals, including Mn (Benedetto et al., 2010; Chakraborty et al., 2015). In fact, overexpressing fpn-1.1 in worms lacking the PD-related gene pdr-1 caused attenuation of Mn toxicity, indicating that ferroportin may be an interesting target against manganism (Chakraborty et al., 2015).
The elucidation of Mn metabolism in C. elegans has also allowed other related genes to be investigated as molecular targets of different tested therapies. Peres et al. (2019) confirmed the therapeutic effect of two small molecules (VU0063088 and VU0026921) that could modify the Mn content in vitro. In this study, it was observed that pre-exposure to the molecules in C. elegans offers protection against Mn-induced dopaminergic neurotoxicity in a smf-2 independent manner (Peres et al., 2019).
Other targets have also been elucidated. The studies indicated that antioxidant systems are important to protect against Mn-induced dopaminergic neurodegeneration, such as the transcription factor skn-1 (orthologous to nrf-2 in mammals) and hsp-1 (homologous to HSP-70) (Avila et al., 2016). Reducing the insulin-like signaling can also protect the dopaminergic neurons from Mn, since knocking out AKT-1/2 and SGK-1 increased the expression of antioxidant genes as sod-3 and gcs-1 and even improved slow basal behavior, which is mainly controlled by the dopaminergic system (Peres et al., 2018).
In addition, a transcriptome-side analysis indicated that genes related to oxidative nucleotide damage, unfolded protein response and innate immunity are the major pathways altered by Mn exposure. Genes for both the antioxidant system (gst-4, glutathione s- transferase), the unfolded protein response (UPR, heat shock proteins, IRE1-mediated UPR) and metallothioneins were significantly increased after 60 and 180min Mn exposure, for example. An increase in the homolog for human egr-1 gene (egrh-1, early growth factor response factor 1) implicated in neurological disorders, such as Alzheimer’s disease and regulation of the cholinergic system was also found (Nicolai et al., 2022).
3.2. Iron
Iron is the most abundant transition metal in the environmental and in the biological systems, involved in biochemical processes that are essential for the maintenance of life in different types of organisms. For instance, Fe is involved in neurotransmitter synthesis, mitochondrial respiration, and oxygen transport among others. Its participation in different metabolic processes occurs mainly due to its capability of electron transfer, as donor and acceptor, assuming two redox states: Ferrous Fe2+ and ferric Fe3+ (Espósito et al., 2020).
However, its essential redox ability is precisely the main mechanism for its neurotoxic effects (Martins et al., 2022). An imbalance in the homeostasis of this metal, as iron overload, can lead to deleterious effects. This essential metal has the ability to catalyze the generation of ROS, causing DNA damage (Klang et al., 2014). Iron overload in the central nervous system (SNC) causes neurotoxicity and cognitive impairments, and has been reported to be involved in the pathogenesis of different disorders, including hypoxic ischemic brain injury and neurodegenerative diseases of aging, such as Alzheimer, Parkinson’s disease, and Huntington disease (Wu et al., 2020).
Fe also has a critical function in brain development and functioning and as aging progresses, ferritin’s ability to store iron diminishes, with the intracellular iron pool increasing (Jenkins et al., 2020). Iron accumulation is common in normal aging, considering the high amounts of iron used by the brain due to its high oxygen consumption. The brain becomes a major producer of reactive ROS, with Fe being one of its main producers. This ROS generation occurs through the Fenton reaction that begins between iron and hydrogen peroxide, generating hydroxyl radical, which is extremely reactive, causing cell damage and neuronal death or ferroptosis by the induction of lipid peroxidation (Abe et al., 2022; Zhou et al., 2020).
The induction of ROS by iron overload also occurs in C. elegans, analogous to the accumulation of Fe with aging (Fagundez et al., 2015; James et al., 2015; Klang et al., 2014). In addition, exposure to Fe may also lead to altered locomotion, characterized as a reduction in the body bends and in the head thrashes. As demonstrated by Hu et al. (2008), exposure to 200μM of iron sulfate causes neuronal alterations in the nematode progeny until the third generation. Authors reported a decrease in the body bends, concomitant with increased rate of chemotaxis, in which the worms did not associate NaCl with the food source and continued to explore the plate, therefore indicating damage to sensory neurons (Hu et al., 2008). Notably, recognition of food source, such as E. coli, is directly mediated by dopaminergic neurons in C. elegans. Indeed, Fagundez et al. (2015) demonstrated neurodegeneration of CEP and ADE dopaminergic neurons caused by 0.05mM of iron sulfate. In this same work, the authors described a decrease in the levamisole induction of egg laying. Degeneration of the cholinergic neurons may cause a defect in egg laying due to the ineffective muscle contraction via cholinergic neurons. The worms also presented a reduction in the motor function and in the mechanic response, strengthen this hypothesis.
As noted before, the increase of Fe in the SNC is a hallmark of neurodegenerative diseases, such as Alzheimer’s and Parkinson’s diseases. James et al. (2015) found increased amount of iron in a model strain for Alzheimer’s disease compared to wild type C. elegans (James et al., 2015). Fe dyshomeostasis has been advanced as one of the hypothesis for oxidative stress and increased total Fe load in the substania nigra in PD patients (Chege and McColl, 2014). Patel et al. (2018) demonstrated that worms expressing alpha-synuclein in the dopaminergic neurons were characterized by neurodegeneration with aging, a process that was partially rescued by treatment with iron chelator deferoxamine. When these worms had deletion of snx-3, a retromer endocytic recycler of ferroportin, they were characterized by increased neurodegeneration, indicating that the exporter is important to control Fe intraneuronal levels. This study also demonstrated that worms with a deletion of smf-3, the ortholog of the divalent metal transporter-3 in humans, have decreased dopaminergic neurodegeneration. Therefore, worms that express α-synuclein in the dopaminergic neurons have an age-dependent degeneration that can be modulated by the protein effectors of Fe transport, reinforcing the concerns of excessive Fe intake (Patel et al., 2018).
3.3. Copper neurotoxicity
Copper (Cu) is the third most abundant essential metal, after iron and zinc (Soares et al., 2017). Essential as a cofactor to a range of enzymatic proteins, mainly by its redox ability—Cu (I) and Cu (II)—it is also necessary for mitochondrial respiration, immune defenses, neurotransmitter synthesis and neurotransmitter regulation.
Indeed, Cu exposure in C. elegans causes neuronal alterations evidenced by reduction in the head thrashes, body bends and ROS production (Liu et al., 2022; Xabier et al., 1992; Zhang et al., 2021c). It is also able of inducing an increase in apoptosis. The exposure to Cu induces the expression cholinergic genes such as ace-1, ace-2, ace-3 and ace-4. ace-1 is involved with the AChE expression in the outer epidermis cells and controls vulval pump, whereas ace-2 is expressed mainly in the neurons. These two genes are responsible for 95% of all AChE activities in C. elegans and these findings correlated with the locomotor alterations observed. There was also an increase in the expression of heat shock proteins hsp-16.1, hsp-16.2, hsp-16.48 that are involved with the control of ROS production, possibly a compensatory response to Cu intoxication (Liu et al., 2022). GABAergic neurodegeneration induced by Cu exposure was also characterized by neuronal loss in nerve cords, as depicted by dorsal and ventral gaps, and reduced fluorescence in the AVL, RMEs and RIS cell body neurons (Du and Wang, 2009).
Zhang et al. (2021b,c) evaluated dopaminergic neurons following exposure to Cu, and showed that all tested concentrations led to decreased fluorescence of dopaminergic neurons and a reduction in dat-1 gene expression. Expression of mtl-1 and mtl-2 were also decreased after exposure to Cu, while sod-3 expression was increased. The authors also evaluated trans-generational effects of Cu for five generations. They observed that a decrease in the body length was observed until the second generation and body bends decrease was observed until the fourth generation (Zhang et al., 2021c). The number of head thrashes was decreased until the second generation and the reduction in the fluorescence intensity in dopaminergic neurons was observed until the third generation at the highest Cu concentration. dop-3 gene expression, responsible for dopamine action in the muscle movements, was decreased at all the concentrations in the parental and F1 and F2 generations, which was consistent with the phenotypic alterations. In addition, Cu exposure changed the expression of genes associated with the regulation and transmission of gene information with metal binding properties. jhdm-1, which expresses the demethylase JHDM-1, and mes-4, which expresses the histone methyltransferase, showed increased expression in the parental generation. mys-1 encodes a histone acetyltransferase, a specific chromatin marker that responds to DNA damage, promoting DNA repair. mRNA level of mys-1 in P0, F1 and F2 was dysregulated following parental Cu exposure, which may result in mis-regulation of DNA repair and dopamine receptor deficiency, ultimately leading to neuronal damage and motor function abnormalities.
3.4. Lead neurotoxicity
The non-essential metal lead (Pb) is an environmental neurotoxicant present in soil, water, automobile exhaust, polluted food, paint, among others. It is classified as the second most toxic metal after arsenic. Total Pb burdens in the human body depend on occupation and levels in the environment. It is presumed that a person with 70 kg will have an average of 120mg of Pb distributed in various tissues. The main route of exposure is through the oral intake, being absorbed in the intestine, transported to the liver, kidney and accumulating in bones and teeth. Exposure to high concentrations of Pb can cause encephalopathies (autism, ADHD) in adults and children. Evidences points out to cognitive decline, alterations in behavior, in brain structure and neurodegenerative processes (Kumar et al., 2020). Children are more susceptible to its neurotoxic effects due to the neuronal developmental stage (Akinyemi et al., 2019; Kumar et al., 2020).
Studies in C. elegans have shown neurotoxic sequalae following Pb early life exposure. Akinyemi et al. (2019) demonstrated dopaminergic neurodegeneration caused by exposure to 2.5 and 5mM of Pb acetate in L1 worms. Pb was shown to decrease the number of body bends, dopamine levels and monoamine oxidase (MAO) activity. mRNA levels of the dat-1 transporter were elevated upon Pb exposure. pkc-1 and pkc-2 mutants exposed to Pb did not show alterations in the number of body bends and in dat-1 expression, suggesting that PKC has an critical role in the dopaminergic neurotoxicity induced by Pb (Akinyemi et al., 2019). Chemotaxis was impaired in early exposed worms as well at concentrations as low as 2.5μM. Morphological alterations, length and a decrease in the fluorescence of the ASE neurons was observed at all the tested concentrations. The ech-3 gene encodes a zinc transcription factor necessary for specification and function of the ASE neurons and exposure to Pb reduced the transcription of this gene. Notably, ech-3 mutant worms also presented a reduction in the chemotaxis assay. The authors suggested that early exposure to Pb interfered with the terminal differentiation of the ASE neurons (Xing et al., 2009a).
Exposure in older worms also led to neurotoxic effects. Exposure for 24 h to 2.5, 75 and 100μM of Pb in L4 worms led to decreased number of head trashes and body bends. These authors analyzed several metals, and Pb was one of the most toxic of all, as it impaired behaviors even at a low concentration of 2.5μM (Wang and Xing, 2008). These results are in agreement with those found by Li et al. (2013a,b), where Pb exposure decreased the number of body bends, head trashes and the reversal frequency, in addition to increasing ROS levels and dopaminergic neurodegeneration (dendrite breaks). This work also demonstrated decreased mRNA levels of ttx-1, tax-2, tax-4 e ceh-14, which are required for the differentiation and function of the AFD neurons (thermosensory neurons) (Li et al., 2013a). Wu et al. (2012b) demonstrated a reduction in the thermotaxis behavior into the concentrations of 2.5, 75, 100 and 150μmol of Pb and a neurodegeneration in the AFD sensory neurons as well. These neurons release neuropeptides and glutamate and therefore, it is possible that glutamatergic neurons are affected by Pb exposure (Wu et al., 2012b).
GABAergic neurotoxicity has been also shown upon late exposure to Pb. Exposure to 2.5, 75 e 200μM for 24 h led to neuronal loss, concomitant with decreased number of GABAergic neurons in head Anterior Ventral Process L (AVL), Ring Motor Neurons E (RMEs) and the Ring Inter Neurons S (RIS) (Du and Wang, 2009). In another study, a 6 h exposure at different larval stages caused GABAergic neuronal loss, increasing the number of gaps in the ventral and dorsal GABAergic neurons. Interestingly, worms exposed at the L4 stage showed higher resistance to the neurotoxic effects when compared to earlier stages, and this resistance was higher in young adults, reinforcing the heightened toxicity of Pb in younger subjects (Xing et al., 2009b). This work also demonstrated neurodegeneration in cholinergic neurons, using resistance to aldicarb and levamisole to evaluate presynaptic and postsynaptic neurons, respectively. The results showed the same profile, with worms exhibiting heightened sensitivity to Pb in the tree first larval stages (Xing et al., 2009b). Overall, the data indicate that Pb is more toxic to young worms at early neurodevelopmental stages.
3.5. Mercury neurotoxicity
Methylmercury (MeHg) is an organomercurial form produced by the methylation of inorganic Hg by bacterium in the environment. MeHg bioaccumulates in the trophic levels, reaching humans mainly by fish consumption. MeHg is lipophilic and exposure to this contaminant can cause severe neurological and cognitive deficits, physical congenital damage, motor and sensory alterations. Occupational exposure to inorganic Hg2+ is involved with disturbance in the central and peripheral system, tremors, uncoordinated movements, neuronal and renal damage (Bianchini et al., 2022; McElwee and Freedman, 2011).
McElwee and Freedman (2011) compared the effects of Hg2+ and MeHg, exposure in C. elegans. Both forms reduced worms’ body length, locomotion, velocity and the mean amplitude of the head movements, but notably MeHg neurotoxicity was more pronounced than Hg2+. In agreement, a chronic exposure to MeHg during 3 days altered the worms locomotion from 1 to 3μg/mL, whereas exposure to inorganic Hg2+ did not alter locomotion activity even at 5μg/mL (Camacho et al., 2022).
The dopaminergic system is one of the targets of this metal. Early life exposure to MeHg causes a reduction in basal slowing response, particularly in pdr-1 mutant worms (homolog of mammalian parkin). These findings indicate a dysfunction of the DAergic neurons after exposure to this toxicant (Martinez-Finley et al., 2013). Indeed, concentrations as low as 0.3–0.5μM of MeHg caused a neurodegeneration in the CEP dopaminergic neurons. Of note, a reduction in the expression of the antioxidant transcription factor SKN-1/Nrf2 increased the vulnerability of the dopaminergic neurons (Vanduyn et al., 2010). Evidence for specific damage to CEP neurons was also observed upon chronic exposure (10 days, 5μM) to MeHg, whereas no alteration in other dopaminergic, cholinergic or glutamatergic neurons were observed (Ke et al., 2020b). A decline in the speed of the swimming behavior in adults exposed for 48 h from L1 stage (0.05–5.0 μM) was also associated with damage to the dopaminergic system. These alterations were suppressed by a cat-2 mutation (homolog of mammalian tyrosine hydroxylase), with suppressed dopamine synthesis (Ke et al., 2021). As a previous study indicated that MeHg induces dopamine release in PC12 cells (Tiernan et al., 2013), it is reasonable to posit that this metal alters the dopamine neuronal metabolism.
Several studies indicated that MeHg toxicity occurs due to its high affinity for thiol groups (-SH). Ke et al. (2020a) reported that the co-treatment for 1 h with N,N′bis-(2-mercaptoethyl) isophthalamide (NBMI) prevented acute MeHg neurotoxicity. NBMI is a lipophilic thiol agent and has high affinity for Hg. The work demonstrated the efficacy of NBMI in protecting DAergic neurons from the toxic effects of chronic exposure to MeHg. Co-treatment with NBMI reduced the formation of puncta by more than 60% and also reduced gst-4 and gcs-1 expression. Of note, mitochondrial function was restored, an important target for increased ROS levels caused by MeHg toxicity (Ke et al., 2020a).
3.6. Cadmium
Cadmium (Cd) is a transition metal considered non-essential because it has no physiological function in living beings (Wang and Du, 2013). In addition to being classified as a toxic metal, it is also carcinogenic (Waalkes, 2000; Wang and Du, 2013). Industrial disposal, battery sources and consumption of tobacco derivatives are sources of Cd exposure, which at high concentrations or in a long-term exposure can bring several health problems, including neurological issues (Mead, 2010).
The CNS is readily affected by Cd exposure. Cd intoxication can lead to headache, vertigo, olfactory loss, parkinsonian symptoms, slow gait, peripheral neuropathy, decreased ability to focus, and learning difficulties (Wang and Du, 2013). Elevated concentrations of Cd in children cause developmental delays, and a reduction in the Intelligence Quotient (IQ) was attributed to the high concentration of Cd (Cao et al., 2009; Mead, 2010; Pihl and Parkes, 1977). Exposure to Cd was associated with damage to the cholinergic, dopaminergic and serotonergic systems, to which the symptoms described above are attributed. Cognitive and emotional dysfunctions described by metal intoxication are related to serotonergic damage (Wang and Du, 2013). Dopaminergic damage was evidenced in early-stage exposure (González-Hunt et al., 2014). Recently, the relationship between serotonergic and reproductive damage on lethality rate after exposure to Cd was demonstrated in the C. elegans model. Worms exposed to Cd showed a reduction in serotonin synthesis by inhibiting the expression of the tryptophan hydroxylase enzyme. Reduction in serotonin in the presynaptic membrane caused impairment in contraction of the vulvar muscles, preventing egg laying and inducing the formation of the bagging phenotype, which occurs by the hatching of the eggs inside the worm, which leads to mortality (Wang et al., 2018).
Oxidative stress is also a hallmark of Cd neurotoxicity. Chronic exposure for 16 h to low concentrations of Cd caused a 25% reduction in the activity of CuZn-SOD (SOD-1) and reduced its expression in C. elegans (Bovio et al., 2021). SOD-1 aggregates have been observed in patients with amyotrophic lateral sclerosis (ALS) (Kaur et al., 2016). Furthermore, alteration in the enzymatic activity of SOD-1 has been described in different neurodegenerative diseases, including PD (de la Torre et al., 1996). Thus, the reduction in SOD-1 enzymatic activity induced by Cd exposure suggests that Cd is an environmental risk factor for neurodegenerative diseases (Bovio et al., 2021).
Stress also occurs through the production of advanced glycation end products (AGEs). AGEs bind to a receptor called RAGE, which has been implicated in hyperglycemia and metal-induced neurotoxicity (Lai et al., 2020). The role of RAGEs in Cd-induced neurotoxicity was determined in C. elegans. Using transgenic GPF strains for RAGEs in dopaminergic and serotonergic neurons, the researchers identified that the co-occurrence of metal exposure and RAGE expression can induce neurodegeneration (Lawes et al., 2020).
3.7. Aluminum
As the most abundant metal in the earth’s crust, Aluminum (Al) is widely found, being present in food and drinking water (Brough and Jouhara, 2020). Although abundant, it does not have any known biological function. It has favorable physico-chemical characteristics that make it suitable for industry (Crisponi et al., 2012b). Among the main sources of exposure to the metal are diet, drinking water, cosmetics and drugs (Tietz et al., 2019).
Despite its wide availability and use, Al is a non-essential metal for human health and can be toxic upon exposure. Various toxic effects are attributed to the metal, including an increased risk of developing breast cancer and bone disorders (Darbre et al., 2013; Klein, 2019). However, these findings still need to be characterized. Its neurotoxic effects are well established. Recent studies linked exposure to Al with different neuropathological conditions, including the development of neurodegenerative diseases, such as Alzheimer’s disease and multiple sclerosis, besides neurodevelopmental deficits and autism spectrum disorder (Exley and Clarkson, 2020; Kumar and Gill, 2009).
Al exposure in C. elegans causes significant developmental delay and disrupts the homeostasis of other metals (Page et al., 2012). Transgenerational effects have also been confirmed in the nematode, where exposure to high concentrations of Al caused deficits in behavioral plasticity, which were transferred to unexposed progeny (Wang et al., 2009a; Ye et al., 2008). In addition, memory deficits were detected by using a thermotaxis assay, a behavior controlled by AIY/AIZ and AIY/RIA cells in a presynaptic location or between AFD/AIY neurons at the post-synaptic location, receiving glutamatergic input. In addition, antioxidant vitamin E post treatment led to recovery of this behavior by increasing calcium neuronal signaling (Ye et al., 2008).
In addition, Al exposure for only 30min causes dopaminergic damage by causing a reduction in ATP levels, altering mitochondrial membrane potential and inducing apoptosis. The damage to the dopaminergic system caused by Al is dependent on the smf-3 metal transporter gene, homolog to mammalian DMT-1 (VanDuyn et al., 2013).
3.8. Metal mixtures
The synergistic effects of metals have been studied to better approximate the real exposure environment to the environment created in the laboratory. In fact, human beings are exposed in different ways to a wide variety of toxic elements, and this is unlikely to occur in isolation.
Mn contamination is an example of common co-exposure, usually coexisting with Cd and Pb in environmental and occupational settings. Studies have demonstrated the effects of these metals in C. elegans in conditions of binary and ternary co-exposure. The results indicated that although the mixture Mn+ Pb presented a synergistic effect on lethality, the dopaminergic damage was similar for both Mn+ Pb and Mn+Cd combinations (Lu et al., 2018). The ternary combination of these metals was also investigated, and the dopaminergic degeneration was like the binary combinations. Authors hypothesized that the potential deactivation of catecholamine-based pathways induced by Mn may counteract the toxic effects from deactivation of serotonin- or acetylcholine-based pathways by Cd, while due to the non-specific nature of Pb toxicity, any binary combination with Pb may potentially increase in effect (Tang et al., 2019). Mn contamination is an example of common co-exposure, usually coexisting with Cd and Pb in environmental and occupational environments. One study demonstrated the effects of these metals on C. elegans in a condition of binary and tertiary co-exposure. In addition to toxicological tests, neurodegenerative effects were evaluated using GFP transgenic strain for dopaminergic neurons, the results indicated a potential neurodegenerative effect for the mixture of Mn, Cd and Pb, presenting a more complex pattern in the analyzed endpoints, in addition to other toxic effects in worms (Tang et al., 2019).
Another mixture that was potentially toxic when compared to their individual exposures was MeHg and Mn. Co-exposure to these metals in L1 worms potentiated neurodegenerative effects by reducing body bends, pharyngeal pumping and increased abnormalities in cholinergic neurons. The expression of acetylcholinesterase ace-2, vesicular monoamine transporter cat-1, and antioxidant genes as sod-3, sod-4 and ctl-3 were increased. Thus, exposure to the MeHg and Mn mixture caused greater cholinergic and monoaminergic damage than exposure to each of the metals alone (Schetinger et al., 2019).
Co-exposure to Zn, Cu and Cd was also evaluated in C. elegans. The behavioral tests demonstrated that Zn has a neutral effect on Cd effects, whereas Cu+Zn or Cu+Cd have additives effects in impairing locomotion, enhancing the damage induced by each metal alone (Moyson et al., 2018).
3.9. Other metals
Recent studies have demonstrated neuronal effects in C. elegans caused by other metals that are not commonly known for their neurotoxicity or that have limited studies about toxicity, such as Arsenic, Cobalt and Nickel.
Arsenic is classified as a highly hazardous substance and is considered WHO’s 10 chemicals of greatest public health concern by the WHO (2022). Being well established as a toxic agent in different tissues, (Zhang et al., 2020), the authors sought to elucidate the transgenerational effects of arsenic intoxication. Worms (F0) were exposed to arsenite at different doses and after analyzing the behavioral effect on the treated worms, the offspring of subsequent generations (F1 and F2) were evaluated. In F0, the data indicated behavioral damage and neuronal degeneration. Behavioral damage was observed in subsequent F1 and F2 generations even absent arsenic exposure. Another study showed that arsenic intoxication was dependent on a bicarbonate transporter, ABTS-1, responsible for regulating cell volume and pH (Liao et al., 2010).
The molecular mechanisms of cobalt intoxication, characterized by cognitive impairment and peripheral neuropathy, were elucidated using C. elegans as an experimental model. In this study, exposure to CoCl2 for 2 h caused mitochondrial fragmentation, growth impairment, apoptosis, autophagy and oxidative stress by activating DRP-1, responsible for mitochondrial fission (Zheng et al., 2020).
The neurotoxic effects of nickel were recently investigated in a model of acute neurotoxicity in C. elegans, after 1 h of exposure to NiCl2 at the first larval stage (L1). Worms showed dose-dependent cholinergic, dopaminergic and GABAergic degeneration. In addition, behavioral assays reflected the effects of Ni on cholinergic and dopaminergic function. The authors confirmed that the observed neuronal effects were associated with oxidative stress by inducing GST-4 expression, suggesting that neuronal damage was induced by oxidative stress due to Ni exposure during worm development (Ijomone et al., 2020).
4. Metallic nanoparticles neurotoxicology in C. elegans
Inorganic nanoparticles are materials with nanometric dimensions with a central core composed of inorganic specimens, especially metals, which give them exciting applications. They have diverse shapes (spherical, layers, oval), different superficial charge that grant stability to the nanostructure, and changes in composition (Mabrouk et al., 2021). NPs have been developed to enable more efficient and low-cost products, seeking better properties and applications. However, the same properties that make nanomaterials attractive may also be responsible for harmful effects in living organisms and concerns about their release into the environment have raised in the last years (Domingues et al., 2022).
Metallic NPs can be composed of metals, metallic oxides and metallic salts (Contado, 2015). As reviewed above, neurons are cells with known vulnerability to metal toxicity; therefore, it is important to observe the effects of metallic NPs in this system (Soares et al., 2017).
Because of their small size, which results in a large surface area, nanoparticles can easily enter the cells from various routes (pulmonary, dermic, gastrointestinal), have increased absorption capacity, cross the blood-brain barrier and remain in the central nervous system for long periods of time (Mushtaq et al., 2015; Win-Shwe and Fujimaki, 2011). NPs can also cause mechanical damage to tissues due to their structure and induce inflammatory processes. Studies have shown that most of the toxicity of various NPs could be attributed to the release of the metallic ions after cellular uptake by phagocytosis. However, there is no conclusive answer so far about whether NP toxicity is due to specific particle effects or the released metal ion or caused by the charge, surface modification, radiation they may emit (Wu et al., 2019).
Oxidative stress has been considered the prime mechanism of NPs causing toxic effects. ROS accumulation have been observed in C. elegans treated with a variety of NPs (Kong et al., 2017; Kumar et al., 2017; Rogers et al., 2015; Wu et al., 2011; Yu et al., 2011; Zhang et al., 2011). The brain is highly susceptible to ROS-induced neurotoxicity due to the high oxygen consumption and rapid metabolic activity (Mao et al., 2010). Likewise, some studies have shown that ROS levels, as reliable indicators of oxidative stress, are closely related to neurotoxicity (Abramov et al., 2007). Therefore, it remains prudent to investigate pathobiological changes of the nervous system in response to NP exposure. Unfortunately, there are few studies assessing the neurotoxicology of metallic NPs in vivo in rodents and in alternative models such as C. elegans.
4.1. Ag-NPs
Ag-NPs are the most studied nanomaterials due to their important properties such as chemical stability, malleability, high electrical and thermal conductivity, catalytic activity and potent antimicrobial action, in addition to relatively low production cost (Durán et al., 2019). The world production of Ag-NPs ranges between 135 and 420 tons per year (Pulit-Prociak and Banach, 2016; Syafiuddin et al., 2017) and their main application are in medical products, clothing, sports equipment, electronics and as food and drinks preservative (Seabra and Durán, 2015; Castellano et al., 2007; Hartemann et al., 2015; Nowack, 2017). Consequently, environmental release of Ag-NPs and its potential neurotoxicity has attracted a great deal of attention (Sørensen and Baun, 2015).
In C. elegans, Ag-NPs accelerate the age-associated decline in swimming and the increase in uncoordinated movements. This behavioral damage induced by NPs was correlated with axonal protein aggregation and neurodegeneration in serotonergic HSN neurons and in sensory ADF neurons, which control egg-laying behavior and locomotor phenotypes, respectively (Seabra and Durán, 2015). C. elegans exposed to Ag-NPs also showed changes in locomotion behaviors such as the number of head thrashes, body bends, pharyngeal pumping frequency and defecation interval, and sensory perception behaviors such as chemotaxis and thermotaxis assays. In addition, a decrease in the fluorescence of dopaminergic, GABAergic and cholinergic neurons was observed in a dose-dependent and time-dependent manner. The same was not observed in glutamatergic neurons, suggesting differential susceptibility of different types of neurons to Ag-NPs exposure in C. elegans. In fact, it has been demonstrated that exposure to low doses of Ag-NPs increased in the gene expression of γ-aminobutyric acid and dopamine receptors in the nematodes. These findings suggest that higher doses of Ag-NPs could cause neurotoxicity through an imbalance in neurotransmitter signaling (Zhang et al., 2021b).
Genetic analysis demonstrated that Ag-NPs exposures consistently reduced several biological processes such as regulation of locomotion, reproduction and cell growth, as well as neuroactive ligand-receptor interaction pathways Wnt and MAPK signaling (Viau et al., 2020). The alterations caused by Ag-NPs exposure in locomotion might be involved with the negative regulation of genes involved in the neuroactive receptor-ligand interaction as well as those of the octopamine receptor family (ser-1) (neuroactive receptor-ligand interaction pathway), skp1 (component of the ubiquitin ligase complex, skr-8, skr-10, skr-12,from the WNT signaling pathway), ver-1 protein (ver-1) and heat shock protein (hsp-70) (both from MAPK signaling pathway) (Viau et al., 2020).
The behavior of the Ag-NPs can be influenced by environmental factors that alter their bioaccumulation and toxicity. The change in the ionic force of the exposure medium in C. elegans (presence of NaCl, phosphate, for instance) significantly increased reproductive toxicity and neurotoxicity (number of head thrashes and body bends) caused by Ag-NPs in C. elegans. In addition, a greater ionic force increased the bioaccumulation of Ag-NPs in E. coli and consequently in nematodes, indicating that the bioavailability and potential ecotoxicity of Ag-NPs are associated with environmental factors (Yang et al., 2018).
Another study using Ag selenide quantum dots (Ag2Se-QDs) demonstrated that these nanomaterials accumulated in the body of nematodes, decreasing lifespan and impairing neurobehaviors such as head thrashes and body bends in C. elegans, a result similar to those found with Ag-NPs (Liang et al., 2022). Overall, these results suggested that Ag-NPs cause neurotoxicity in C. elegans.
4.2. Fe-NPs
Fe nanoparticles possess a high magnetic nature, high surface area, electrical and thermal conductivity. They also have dimensional stability. Due to the properties of Fe nanoparticles, they are known as magnetic nanoparticles. Magnetic nanoparticles have been explored widely in the decades due to their large number of applications in the field of spintronics, biology, and medical science (Umair et al., 2016). Nanoparticles that are made of ferro-or ferromagnetic materials below a certain size (generally 10–20 nm) can exhibit a unique form of magnetism called superparamagnetism. Although the potential benefits of Fe-NPs are considerable, exposure can result in significant cytotoxicity. When cells are exposed to high doses of Fe oxide nanoparticles, excess ROS formation occurs, affecting the normal functioning of the cells, leading to apoptosis or cell death, growth delay, development abnormality and decreased fertility (Valdiglesias et al., 2016).
The impact of surface charge on polymeric shell coated FeO-NPs has been studied in C. elegans. When coated with the negatively charged polymer PMDA (Poly-Maleic Acid Conjugated with Dodecylamine), these NPs have been proven to be toxic, shortening the lifespan of the nematode, increasing ROS production and reducing the rate of pharyngeal pumping. Of note, decreased pharyngeal pumping is an indicative of the presence of neurotoxicants. On the other hand, zwitterionic NPs (partially cationic and anionic) exhibited lower toxicity, better distribution, and higher Fe delivery (Amigoni et al., 2021). In another study, DMSA (Dimercaptosuccinic acid) coated FeO-NPs depicted toxicity by altering growth, reproduction, locomotion, pharyngeal pumping, defection and autofluorescence (Wu et al., 2012a). Other forms of Fe-NPs such as cube-like Fe nitride (α”-Fe16N2) and zero valent Fe NPs (nZVI) showed lower or no neurotoxicity in the nematodes, although oxidative stress was found in exposed animals (Gubert et al., 2022; Yang et al., 2016). FeO-NPs have utility in a wide variety of applications. While their benefits are considerable and they are considered of low risk, their neurotoxic potential must be further investigated (Karlsson et al., 2009; Singh et al., 2010). Even the coating strategy was not effective in reducing their toxicity, and therefore much work needs to be done to reduce their potential impact on living organisms.
4.3. Al-NPs
Aluminum oxide nanoparticles are increasingly used in domestic, industrial and medicine scales (M’rad et al., 2018). Aluminum (Al) is a vital etiopathogenic agent and has been associated with the incidence of neurodegenerative diseases (Halatek et al., 2005). However, despite the fact that some nanoparticles are shown to cause neurotoxic effects in rodents, their influence on the nervous system is limited (M’rad et al., 2018). Studies that investigate the toxicity of Al-NPs indicate that, as most of NPs, toxicity occurs by oxidative stress.
More recently, the acute toxicity of Al2O3 NP- has been examined in C. elegans with lethality, growth, reproduction and stress response outcomes (Wang et al., 2009b; Wu et al., 2011). Nematodes chronically exposed to Al2O3-NPs for 10 days suffered abnormal locomotion behavior, which was associated with increased ROS production and suppressed antioxidant defenses. The antioxidant enzymes encoded by the sod-2 and sod-3 genes were down-regulated and the mutants were more susceptible to these NPs, implying that compromised antioxidant mechanisms may also contribute to their neurotoxic effects (Li et al., 2012b). After exposure for 3 days, significant neurotoxicity was noted characterized by decreased number of head thrashes, in addition to decreased locomotion speed, which were more significant for Ag- and Si-NPs (Viau et al., 2020). When compared to Si-, Ti- and polystyrene NPs, Al2O3-NPs was the most neurotoxic (Li et al., 2020).
Li et al. (2013a,b) demonstrated that glutamate transporter EAT-4, serotonin transporter MOD-5, and dopamine transporter DAT-1 might pose important molecular targets of Al2O3-NPs neurotoxicity. They observed that the hypo-locomotor effect is driven by movement inhibition of downstream effectors D1-like dopamine receptor DOP-1, ionotropic serotonin receptor MOD-1, and non-NMDA glutamate receptors GLR-2 and GLR-6 (Li et al., 2013b).
Surface modifications were assessed with Al2O3– NPs in a recent study, which compared pristine Al2O3-NPs (p-Al2O3), hydrophilic (w-Al2O3) and lipophilic (o-Al2O3) modifications (Zhang et al., 2021a). The lipophilic form caused higher toxicity by reducing head trashes, increasing ROS formation and mitochondrial damage and inducing apoptosis. Transcriptomic analysis indicated not only differential expression in genes related to glutathione metabolism and peroxisome pathways, but also in genes that control locomotion, since the exposure to this metal changes the expression of several genes responsible for various biological processes in the worms (Zhang et al., 2021a).
4.4. Cu-NPs
Copper toxicity is widely known, mainly due to the intracellular accumulation that leads to disturbances of mitochondrial metabolic enzymes and cell death (Kahison and Dixon, 2022). Thus, studies evaluating the neurotoxicity of Cu-NPs deserve attention, since these NPs have considerable cytotoxicity and can cause long-term effects (Fahmy and Cormier, 2009; Ivask et al., 2010; Keller et al., 2013). Currently, CuO-NPs are mainly used in electronics, including gas sensors, batteries and solar energy, as these NPs have great superconducting properties. In addition, CuO-NPs are also incorporated into various personal and household hygiene products due to their bactericidal action (Bondarenko et al., 2013).
Exposure to CuO-NPs has been shown to cause toxic effects on behavioral assays by reducing locomotion velocity, head thrashes and pharyngeal pumping in C. elegans (Viau et al., 2020). The study suggests that Cu-free ions released by the NPs in the solution may have been absorbed by the worm’s food source, E. coli. After absorption, Cu ions can be found in the head and the worms’ whole body (Gao et al., 2008). Eating behavior was considered the most sensitive toxicological endpoint after CuO-NPs exposure since it was affected by all tested concentrations (Viau et al., 2020). Notably, intestinal intake of CuO-NPs is probably the entry point of NPs in C. elegans, considering that the cuticle is poorly permeable (Mashock et al., 2016; Song et al., 2014). Data show that this reduction in the nematode eating behavior is a response to environmental stress (through sensory neurotransmitters) and changes in pharyngeal pumping activity (Raizen et al., 2018). In addition, high Cu levels have been reported to induce paralysis, reducing feeding and reproduction by affecting pharyngeal pumping and egg laying, respectively (Peterson et al., 2008). These behavioral alterations may be related to the degeneration of dopaminergic neurons. Furthermore, mutants of the bivalent metal transporters, smf-1 or smf-2, showed greater tolerance to Cu exposure, implicating both transporters in Cu-induced neurodegeneration (Mashock et al., 2016).
4.5. Au-NPs
Gold nanoparticles (AuNPs) are a biocompatible class of nanomaterials widely used for bioimaging (Chen et al., 2015a; Shang et al., 2013), biosensing (Chen and Tseng, 2012; Liu et al., 2015), facial creams (Guix et al., 2008), and targeted therapeutic purposes (Wang et al., 2011; Zhang et al., 2014). Nevertheless, these nanoparticles can be directly taken up by animals, affecting their physiology and thus posing a toxicological risk. The uptake and toxicity of AuNPs highly depends on their shape ( Jin et al., 2013), size (Ma et al., 2011) and surface charge (Albanese et al., 2012; Jiang et al., 2015).
Hu et al. (2018a) demonstrated significant changes in worm locomotion after exposure to AuNPs, raising the hypothesis that this effect is based on impaired motor neuron function. Therefore, when examining primary neurons isolated from C. elegans, before and after exposure to AuNPs, it was found that axonal growth was significantly disrupted by exposure to AuNPs (Hu et al., 2018a). Neuronal development impairment may be related to altered expression of cut-3 and fil-1 (both involved in body morphogenesis) and dpy-14 (expressed in embryonic neurons), in addition to mtl-1, important for metal detoxification and homeostasis (Mo et al., 2020).
A toxicogenomic analysis demonstrated that AuNPs are bioavailable and cause adverse effects to C. elegans by activating both general and specific biological pathways. There were a total of six upregulated C. elegans genes (apl-1, clp-1, kin-19, par-1, kin-1 and kin-2) homologous to the human amyloid processing pathways (Selkoe, 2001; Tsyusko et al., 2012). The up-regulation of apl-1, amyloid precursor like protein, which is a homolog of the human amyloid precursor protein (app-1) is associated with Alzheimer’s disease in humans (Zheng and Koo, 2011), was confirmed with qRT-PCR. In C. elegans, apl-1 is expressed in multiple tissues, being important for molting, morphogenesis, and larval survival (Hornsten, 2001). Together with Feh-1 (homologue of mammalian Fe65), which is expressed in neuromuscular structure of the pharynx and which was also up-regulated in response to AuNPs, apl-1 forms a complex, controlling the rate of pharyngeal pumping (Zambrano et al., 2002). Even though C. elegans does not accumulate β-amyloid, activation of genes related to the amyloid processing pathways may be indicative of events leading to neurodegeneration and can be a result of an unfolded protein response activated in response to Au NP exposure (Tsyusko et al., 2012).
4.6. Ti-NPs
TiO2-NPs are the most studied metal oxide NPs today because they are widely used as pigments and additives for paints, papers, ceramics, plastics, foods and pharmaceuticals (Aitken et al., 2006; Gélis et al., 2003). Therefore, these NPs have become an important environmental contaminant and studies have been carried out to characterize their neurotoxicity.
Exposure to TiO2-NPs suspensions induces lethality in C. elegans, in addition to significantly reducing body length, head movement and body curvature of these animals (Li et al., 2012a, 2020; Viau et al., 2020; Wu et al., 2012c, 2013). Furthermore, sod-2, sod-3, mtl-2 and hsp-6.48 expression were altered by exposure and associated to the neurotoxicity of TiO2-NPs, as locomotion behavior was impaired (Wu et al., 2014). Dong et al. (2018) investigated the combined effects of TiO2-NPs and nanopolystyrene at environmentally relevant concentrations. Co-exposure to nanopolystyrene particles enhanced the toxicity of TiO2-NPs by decreasing locomotion behavior and inducing intestinal ROS levels in both WT nematodes and sod-3 mutants at higher levels than TiO2-NPs alone (Dong et al., 2018).
Prolonged exposure to TiO2-NPs also decreased the fluorescence intensity of AVL and DVB GABAergic neurons, a degeneration that could not be recovered (Zhao et al., 2014). C. elegans neurons were able to directly uptake anatase and rutile NPs (mineral forms of TiO2) and this uptake inhibited axonal growth, associated with a decline in body length and worm locomotion (Hu et al., 2018b, 2020). Changes in genes related to metal binding or metal detoxification (mtl-2, C45B2.2 and nhr-247), genes involved in worm growth and body morphogenesis (mtl-2, F26F2.3, C38C3.7 and nhr-247), and genes involved in neuronal function (C41G6.13, C45B2.2, srr-6, K08C9.7 and C38C3.7) were also observed. In addition, cl-1 (stress resistance regulator), wah-1 (oxidoreductase), gst-3 (glutathione-S-transferase) and cyp-33C1 (cytochrome P450) were also altered, suggesting a relationship between neurotoxicity and the increase in ROS levels (Hu et al., 2018b).
4.7. Cd-NPs
The transition heavy metal cadmium is a ubiquitous toxic substance present in soil, water, air and food. As a consequence of anthropic activity, the environmental concentrations of Cd increased steadily in many areas around the world (Faroon et al., 2013). Cd is readily absorbed and accumulated in different tissues and organs. Exposure to Cd may have a potential impact on human health and wildlife (Davison et al., 1988).
In nanotechnology, Cd is primarily utilized in the construction of particles known as quantum dots (QDs), which are semiconductor metalloid-crystal structures of approximately 2–100 nm, containing about 200–10.000 atoms (Juzenas et al., 2008; Smith et al., 2008). Due to their small size, QDs have unique optical and electronic properties that impart the nanoparticle with a bright, highly stable, “size-tunable” fluorescence. The large surface area imparted by small size also makes QDs readily functionalized with targeting ligands for site-directed activity. Based on these properties, QDs have the potential for revolutionizing biological imaging at the cellular level, cancer detection and treatment, radio- and chemo sensitizing agents, and targeted drug delivery (Alivisatos, 2004; Hardman, 2006; Juzenas et al., 2008; Smith et al., 2008). However, enthusiasm for QDs is somewhat diluted by the fact that QDs contain substantial amounts of cadmium in a highly reactive form (Rzigalinski and Strobl, 2009).
Exposure to CdSe/ZnS quantum dots has been documented to inflict a defective egg laying phenotype in C. elegans, which implies that motor neurons are involved in reproduction (Galdiero et al., 2020). RME neurons of the head region and AVL and DVB neurons that control defection showed developmental and functional defects after exposure to CdTe quantum dots at the L1 stage of the worm (Zhao et al., 2015). Wu et al. (2015) demonstrated that CdTe QDs alter locomotion, pharyngeal pumping, defecation cycle and even plasticity in learning and memory, time and dose-dependently. At the molecular level, the biological endpoints were likely to be associated with altered expression of genes involved in glutamate, serotonin and dopamine signaling coupled with markedly increased ROS generation (Wu et al., 2015). The pathway by which QDs inflict neurotoxicity in C. elegans involves deposition inside the neurons, perturbations of neuronal development and abnormal neurotransmission. Depending on the experimental conditions, the response seems to be mediated by various degrees of metal toxicity from Cd2+ leaching out of the core in conjunction with NP derived oxidative stress (Sinis et al., 2019).
4.8. Si-NPs
Due to the unique properties of silica NPs such as surface functionality, controllable particle size, and desirable biocompatibility, studies have explored these nanomaterials (Shahbazi et al., 2012; Tao et al., 2014; Zhang and Kong, 2015). The large-scale production and application of Si-NPs increased the risk of human exposure, making it necessary to investigate their neurotoxic effects (Jiang and Gao, 2016).
Prolonged exposure to a form of Si-NPs, mSi-NPs (mesoporous silica nanoparticles), showed locomotion degeneration, shrinkage behavior, and abnormal foraging behavior, which were associated with deficits in the development of GABAergic neurons, including D-type and RME motor neurons. In addition, there was excessive generation of ROS contributing to neuronal damage (Liang et al., 2020b).
Exposure to SiO2-NPs led to toxic effects on locomotion velocity and head thrashes, in addition to inducing ROS production in C. elegans (Li et al., 2020; Wu et al., 2013).
Genetic analysis demonstrated that SiO2-NPs and Si-NPs exposures consistently reduced several biological processes such as regulation of locomotion, reproduction and cell growth (Liang et al., 2020a), as well as neuroactive ligand-receptor interaction pathways, Wnt and MAPK signaling, involved in the regulation of cell reproduction and growth (Viau et al., 2020).
Nematodes exposed to Si-NPs significantly accumulated insoluble and ubiquitinated proteins, which did not occur with worms exposed to bulk silica. Ubiquitin is a protein that marks malformed proteins to be degraded. An analysis of gut cells showed the formation of amyloid-like structures in the worms treated with Si-NPs. These results indicate that Si-NPs induce fibrillation of endogenous proteins and amyloid aggregates. An increase in insoluble proteins in worms treated with Si-NPs was observed throughout their life, altering protein homeostasis in aged worms (Scharf et al., 2013). Pharyngeal pumping rate was also significantly reduced, an endpoint that indicates accelerated aging and muscle atrophy from altered proteostasis (Scharf et al., 2013). Analysis of serotonergic neural cells revealed that protein aggregation induced by Si-NPs was evident in axons of HSN neurons, with presynaptic accumulation of serotonin, inducing disturbed axonal transport that reduces neurotransmission capacity and egg laying of C. elegans (Scharf et al., 2016). As insoluble proteins are involved in the aggregation of the beta-amyloid peptide in Alzheimer’s disease (AD) and alpha-synuclein in Parkinson’s disease (PD) (Pereira et al., 2022), exposure to Si-NPs may cause long-term neurotoxicity by insoluble proteins formation.
4.9. Other metallic nanoparticles
Studies on metallic nanoparticles neurotoxicity are scarce, particularly of some metals that will be described below.
CeO2-NPs are mainly used as catalysts (Prasad and Rattan, 2010), diesel fuel additive (Park et al., 2008) and for glass and ceramic applications (Younis et al., 2016). However, CeO2 NPs have been considered one of the most toxic NPs to the environment. Exposure of CeO2-NPs caused an impairment in head thrashes in C. elegans, but the mechanism of this toxicity is still unclear and needs to be studied (Viau et al., 2020). CeO2-NPs cause an increase in ROS levels in C. elegans, and this imbalance in redox homeostasis may be related to the observed behavioral effects (Zhang et al., 2011).
ZnO-NPs have a wide range of applications, including antimicrobial properties, widely used by the cosmetics and paint industries ( Jiang et al., 2018). In addition, they are used in agriculture to control foodborne pathogens (Nohynek et al., 2010; Tayel et al., 2011). Despite their commercial advantages, studies have shown that exposure to ZnO-NPs may cause toxicity, mainly due to the accumulation in the environment and increased production of ROS (Nowack et al., 2011; Schilling et al., 2010). ZnO-NPs have been shown to cause harmful effects on C. elegans locomotion (Huang et al., 2017; Wu et al., 2013). This behavioral change was associated with decreased ATP levels and significant increase in intracellular ROS levels and lipid peroxidation, suggesting mitochondrial damage, triggering neuronal death (Huang et al., 2017, 2019). Furthermore, ZnO-NPs triggered the translocation of the transcription factor DAF-16/FOXO from the cytoplasm to the nucleus and activated the expression of stress-responsive genes mtl-1 and sod-3 (Huang et al., 2019). Therefore, oxidative stress appears to be strongly associated to the neurotoxicity caused by these NPs.
5. Perspectives and concluding remarks
Using the alternative animal model C. elegans to investigate and understand the neurotoxicity caused by metals and metallic NPs has been fruitful and important for human health and environmental impact assessments. As discussed here, several neuronal systems can be affected, as evidenced by behavioral assays and by in vivo visualization though GFP-tagged neuronal proteins. The main systems affected by metals are DAergic, serotonergic, GABAergic and glutamatergic (sensory neurons). Furthermore, transcriptomic analysis and the use of mutants has indicated that metals and NPs cause neurotoxic effects by disrupting the neurotransmitter signaling (mainly neurotransmitter synthesis and degradation/reuptake processes), by altering the antioxidant enzymatic response, by inducing apoptosis and by interfering in cell division and differentiation. It is concerning that some studies have evidenced the ability of some metals and NPs to promote epigenetic transgenerational inheritance neurotoxicity, which should be further investigated for all metals and NPs.
It is noteworthy some limitations of the model and of the protocols used. It is important to investigate the effects of metals and metallic NPs in all worms stages and using a wide range of concentrations and times of exposure. However, the exposure protocols are so different among groups that makes hard to compare the levels of neurotoxicity. Furthermore, the exposure medium may differ among studies, which can totally change the toxicological outcomes. C. elegans also has limitations for neurotoxicological assessments. It does not have a cardiovascular system, no brain, liver or kidneys. On the other hand, the model respects the 3R policy (reduce, refine and replace), which are pillars of the modern toxicology to avoid the unnecessary use of vertebrates. Finally, the findings from C. elegans have been providing information to help understand potential risks of metals and metallic NPs, especially at the environmental level and to be extrapolated to human outcomes.
References
- Abe C, Miyazawa T, Miyazawa T, 2022. Current use of fenton reaction in drugs and food. Molecules 27, 5451. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Abramov AY, Scorziello A, Duchen MR, 2007. Three distinct mechanisms generate oxygen free radicals in neurons and contribute to cell death during anoxia and reoxygenation. J. Neurosci. 27, 1129–1138. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Aitken RJ, Chaudhry MQ, Boxall AB, Hull M, 2006. Manufacture and use of nanomaterials: current status in the UK and global trends. Occup. Med. (Lond) 56, 300–306. [DOI] [PubMed] [Google Scholar]
- Akinyemi AJ, Miah MR, Ijomone OM, Tsatsakis A, Soares FAA, Tinkov AA, Skalny AV, Venkataramani V, Aschner M, 2019. Lead (Pb) exposure induces dopaminergic neurotoxicity in Caenorhabditis elegans: involvement of the dopamine transporter. Toxicol. Rep. 6, 833–840. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Albanese A, Tang PS, Chan WC, 2012. The effect of nanoparticle size, shape, and surface chemistry on biological systems. Annu. Rev. Biomed. Eng. 14, 1–16. [DOI] [PubMed] [Google Scholar]
- Alivisatos P, 2004. The use of nanocrystals in biological detection. Nat. Biotechnol. 22, 47–52. [DOI] [PubMed] [Google Scholar]
- Amigoni L, Salvioni L, Sciandrone B, Giustra M, Pacini C, Tortora P, Prosperi D, Colombo M, Regonesi ME, 2021. Impact of tuning the surface charge distribution on colloidal iron oxide nanoparticle toxicity investigated in Caenorhabditis elegans. Nanomaterials (Basel) 11, 934–975. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Avila D, Helmcke K, Aschner M, 2012. The Caenorhabiditis elegans model as a reliable tool in neurotoxicology. Hum. Exp. Toxicol. 31, 236–243. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Avila DS, Benedetto A, Au C, Bornhorst J, Aschner M, 2016. Involvement of heat shock proteins on Mn-induced toxicity in Caenorhabditis elegans. BMC Pharmacol. Toxicol. 17, 54. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Balachandran RC, Mukhopadhyay S, McBride D, Veevers J, Harrison FE, Aschner M, Haynes EN, Bowman AB, 2020. Brain manganese and the balance between essential roles and neurotoxicity. J. Biol. Chem. 295, 6312–6329. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Balali-Mood M, Naseri K, Tahergorabi Z, Khazdair MR, Sadeghi M, 2021. Toxic mechanisms of five heavy metals: mercury, lead, chromium, cadmium, and arsenic. Front. Pharmacol. 12, 643972. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Banks WA, 2016. From blood–brain barrier to blood–brain interface: new opportunities for CNS drug delivery. Nat. Rev. Drug Discov. 15, 275–292. [DOI] [PubMed] [Google Scholar]
- Bargmann CI, 1998. Neurobiology of the Caenorhabditis elegans genome. Science 282, 2028–2033. [DOI] [PubMed] [Google Scholar]
- Benedetto A, Au C, Avila DS, Milatovic D, Aschner M, 2010. Extracellular dopamine potentiates mn-induced oxidative stress, lifespan reduction, and dopaminergic neurodegeneration in a BLI-3-dependent manner in Caenorhabditis elegans. PLoS Genet. 6, e1001084. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bianchini MC, Soares LFW Jr., Sousa J, Ramborger BP, Gayer MC, Bridi JC, Roehrs R, Pinton S, Aschner M, Ávila DS, Puntel RL, 2022. MeHg exposure impairs both the catecholaminergic and cholinergic systems resulting in motor and non-motor behavioral changes in Drosophila melanogaster. Chem. Biol. Interact. 365, 110121. [DOI] [PubMed] [Google Scholar]
- Bondarenko O, Juganson K, Ivask A, Kasemets K, Mortimer M, Kahru A, 2013. Toxicity of Ag, CuO and ZnO nanoparticles to selected environmentally relevant test organisms and mammalian cells in vitro: a critical review. Arch. Toxicol. 87, 1181–1200. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bornhorst J, Chakraborty S, Meyer S, Lohren H, Brinkhaus SG, Knight AL, Caldwell KA, Caldwell GA, Karst U, Schwerdtle T, Bowman A, Aschner M, 2014. The effects of pdr1, djr1.1 and pink1 loss in manganese-induced toxicity and the role of α-synuclein in C. elegans. Metallomics 6, 476–490. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bovio F, Sciandrone B, Urani C, Fusi P, Forcella M, Regonesi ME, 2021. Superoxide dismutase 1 (SOD1) and cadmium: a three models approach to the comprehension of its neurotoxic effects. Neurotoxicology 84, 125–135. [DOI] [PubMed] [Google Scholar]
- Bowler RM, Kornblith ES, Gocheva VV, Colledge MA, Bollweg G, Kim Y, Beseler CL, Wright CW, Adams SW, Lobdell DT, 2015. Environmental exposure to manganese in air: associations with cognitive functions. Neurotoxicology 49, 139–148. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Brenner S, 1973. The genetics of behaviour. Br. Med. Bull. 29, 269–271. [DOI] [PubMed] [Google Scholar]
- Brough D, Jouhara H, 2020. The aluminium industry: a review on state-of-the-art technologies, environmental impacts and possibilities for waste heat recovery. International Journal of Thermofluids 1, 100007. [Google Scholar]
- Caito S, Aschner M, 2015. Neurotoxicity of metals. Handb. Clin. Neurol. 131, 169–189. [DOI] [PubMed] [Google Scholar]
- Camacho J, de Conti A, Pogribny IP, Sprando RL, Hunt PR, 2022. Assessment of the effects of organic vs. inorganic arsenic and mercury in Caenorhabditis elegans. Curr. Res. Toxicol. 3, 100071. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cao Y, Chen A, Radcliffe J, Dietrich KN, Jones RL, Caldwell K, Rogan WJ, 2009. Postnatal cadmium exposure, neurodevelopment, and blood pressure in children at 2, 5, and 7 years of age. Environ. Health Perspect. 117, 1580–1586. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Castellano JJ, Shafii SM, Ko F, Donate G, Wright TE, Mannari RJ, Payne WG, Smith DJ, Robson MC, 2007. Comparative evaluation of silver-containing antimicrobial dressings and drugs. Int. Wound J. 4, 114–122. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chakraborty S, Chen P, Bornhorst J, Schwerdtle T, Schumacher F, Kleuser B, Bowman AB, Aschner M, 2015. Loss of pdr-1/parkin influences Mn homeostasis through altered ferroportin expression in C. elegans. Metallomics 7, 847–856. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chalfie M, Tu Y, Euskirchen G, Ward WW, Prasher DC, 1994. Green fluorescent protein as a marker for gene expression. Science 263, 802–805. [DOI] [PubMed] [Google Scholar]
- Chege PM, McColl G, 2014. Caenorhabditis elegans: a model to investigate oxidative stress and metal dyshomeostasis in Parkinson’s disease. Front. Aging Neurosci. 6, 89. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chen TH, Tseng WL, 2012. (Lysozyme type VI)-stabilized Au8 clusters: synthesis mechanism and application for sensing of glutathione in a single drop of blood. Small 8, 1912–1919. [DOI] [PubMed] [Google Scholar]
- Chen L-Y, Wang C-W, Yuan Z, Chang H-T, 2015a. Fluorescent gold nanoclusters: recent advances in sensing and imaging. Anal. Chem. 87, 216–229. [DOI] [PubMed] [Google Scholar]
- Chen P, Chakraborty S, Peres TV, Bowman AB, Aschner M, 2015b. Manganese-induced neurotoxicity: from C. elegans to humans. Toxicol. Res. (Camb) 4, 191–202. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chen P, DeWitt MR, Bornhorst J, Soares FA, Mukhopadhyay S, Bowman AB, Aschner M, 2015c. Age- and manganese-dependent modulation of dopaminergic phenotypes in a C. elegans DJ-1 genetic model of Parkinson’s disease. Metallomics 7, 289–298. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Contado C, 2015. Nanomaterials in consumer products: a challenging analytical problem. Front. Chem. 3, 48. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Crisponi G, Nurchi VM, Bertolasi V, Remelli M, Faa G, 2012a. Chelating agents for human diseases related to aluminium overload. Coord. Chem. Rev. 256, 89–104. [Google Scholar]
- Crisponi G, Nurchi VM, Crespo-Alonso M, Toso L, 2012b. Chelating agents for metal intoxication. Curr. Med. Chem. 19, 2794–2815. [DOI] [PubMed] [Google Scholar]
- Darbre PD, Mannello F, Exley C, 2013. Aluminium and breast cancer: sources of exposure, tissue measurements and mechanisms of toxicological actions on breast biology. J. Inorg. Biochem. 128, 257–261. [DOI] [PubMed] [Google Scholar]
- Davison A, Taylor AN, Darbyshire J, Chettle D, Guthrie C, O’Malley D, Mason H, Fayers P, Venables K, Pickering C, 1988. Cadmium fume inhalation and emphysema. Lancet 331, 663–667. [DOI] [PubMed] [Google Scholar]
- de la Torre MR, Casado A, Lopez-Fernandez ME, Carrascosa D, Casado MC, Venarucci D, Venarucci V, 1996. Human aging brain disorders: role of antioxidant enzymes. Neurochem. Res. 21, 885–888. [DOI] [PubMed] [Google Scholar]
- Domingues C, Santos A, Alvarez-Lorenzo C, Concheiro A, Jarak I, Veiga F, Barbosa I, Dourado M, Figueiras A, 2022. Where is nano today and where is it headed? A review of nanomedicine and the dilemma of nanotoxicology. ACS Nano 16, 9994–10041. [DOI] [PubMed] [Google Scholar]
- Dong S, Qu M, Rui Q, Wang D, 2018. Combinational effect of titanium dioxide nanoparticles and nanopolystyrene particles at environmentally relevant concentrations on nematode Caenorhabditis elegans. Ecotoxicol. Environ. Saf. 161, 444–450. [DOI] [PubMed] [Google Scholar]
- Du M, Wang D, 2009. The neurotoxic effects of heavy metal exposure on GABAergic nervous system in nematode Caenorhabditis elegans. Environ. Toxicol. Pharmacol. 27, 314–320. [DOI] [PubMed] [Google Scholar]
- Durán N, Rolim WR, Durán M, Fávaro WJ, Seabra AB, 2019. Nanotoxicology of silver nanoparticles: toxicity in aninals and humans. Qúım. Nova 42, 206–213. [Google Scholar]
- Espósito BP., Martins AC., de Carvalho RRV., Aschner M., 2020. High throughput fluorimetric assessment of iron traffic and chelation in iron-overloaded Caenorhabditis elegans. Biometals 33, 255–267. [DOI] [PubMed] [Google Scholar]
- Exley C, Clarkson E, 2020. Aluminium in human brain tissue from donors without neurodegenerative disease: a comparison with Alzheimer’s disease, multiple sclerosis and autism. Sci. Rep. 10, 7770. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Fagundez D.d.A., Câmara DF, Salgueiro WG, Noremberg S, Luiz Puntel R, Piccoli JE, Garcia SC, Da Rocha JBT, Aschner M, Ávila DS, 2015. Behavioral and dopaminergic damage induced by acute iron toxicity in Caenorhabditis elegans. Toxicol. Res. 4, 878–884. [Google Scholar]
- Fahmy B, Cormier SA, 2009. Copper oxide nanoparticles induce oxidative stress and cytotoxicity in airway epithelial cells. Toxicol. In Vitro 23, 1365–1371. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Faroon O, Ashizawa A, Wright S, Tucker P, Jenkins K, Ingerman L, Rudisill C, 2013. Toxicological Profile for Cadmium. Agency for Toxic Substances and Disease Registry (US), Atlanta, GA. [PubMed] [Google Scholar]
- Galdiero E, Siciliano A, Lombardi L, Falanga A, Galdiero S, Martucci F, Guida M, 2020. Quantum dots functionalized with gH625 attenuate QDs oxidative stress and lethality in Caenorhabditis elegans: a model system. Ecotoxicology 29, 156–162. [DOI] [PubMed] [Google Scholar]
- Gao Y, Liu N, Chen C, Luo Y, Li Y, Zhang Z, Zhao Y, Zhao B, Iida A, Chai Z, 2008. Mapping technique for biodistribution of elements in a model organism, Caenorhabditis elegans, after exposure to copper nanoparticles with microbeam synchrotron radiation X-ray fluorescence. J. Anal. At. Spectrom 23, 1121–1124. [Google Scholar]
- G!elis C., Girard S., Mavon A., Delverdier M., Paillous N., Vicendo P., 2003. Assessment of the skin photoprotective capacities of an organo-mineral broad-spectrum sunblock on two ex vivo skin models. Photodermatol. Photoimmunol. Photomed. 19, 242–253. [DOI] [PubMed] [Google Scholar]
- Gitler AD, Chesi A, Geddie ML, Strathearn KE, Hamamichi S, Hill KJ, Caldwell KA, Caldwell GA, Cooper AA, Rochet JC, Lindquist S, 2009. Alpha-synuclein is part of a diverse and highly conserved interaction network that includes PARK9 and manganese toxicity. Nat. Genet. 41, 308–315. [DOI] [PMC free article] [PubMed] [Google Scholar]
- González-Hunt CP, Leung MC, Bodhicharla RK, McKeever MG, Arrant AE, Margillo KM, Ryde IT, Cyr DD, Kosmaczewski SG, Hammarlund M, Meyer JN, 2014. Exposure to mitochondrial genotoxins and dopaminergic neurodegeneration in Caenorhabditis elegans. PloS One 9, e114459. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Goodman MB, Hall DH, Avery L, Lockery SR, 1998. Active currents regulate sensitivity and dynamic range in C. elegans neurons. Neuron 20, 763–772. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gubert P, Puntel B, Lehmen T, Bornhorst J, Avila DS, Aschner M, Soares FAA, 2016. Reversible reprotoxic effects of manganese through DAF-16 transcription factor activation and vitellogenin downregulation in Caenorhabditis elegans. Life Sci. 151, 218–223. [DOI] [PubMed] [Google Scholar]
- Gubert P, Puntel B, Lehmen T, Fessel JP, Cheng P, Bornhorst J, Trindade LS, Avila DS, Aschner M, Soares FAA, 2018. Metabolic effects of manganese in the nematode Caenorhabditis elegans through DAergic pathway and transcription factors activation. Neurotoxicology 67, 65–72. [DOI] [PubMed] [Google Scholar]
- Gubert G, Gubert P, Sandes JM, Bornhorst J, Alves LC, Quines CB, Mosca DH, 2022. The nanotoxicity assessment of cube-like iron nitride magnetic nanoparticles at the organismal level of nematode Caenorhabditis elegans. Nanotoxicology 16, 472–483. [DOI] [PubMed] [Google Scholar]
- Guix M, Carbonell C, Comenge J, GarcÍa-Fernández L, Alarcón A, Casals E, 2008. Nanoparticles for cosmetics: how safe is safe? Contrib. Sci. 4, 213–217. [Google Scholar]
- Gupta S, Dhanda S, Sandhir R, 2019. Anatomy and Physiology of Blood-Brain Barrier, Brain Targeted Drug Delivery System. Elsevier, pp. 7–31. [Google Scholar]
- Halatek T, Sinczuk-Walczak H, Rydzynski K, 2005. Prognostic significance of low serum levels of Clara cell phospholipid-binding protein in occupational aluminium neurotoxicity. J. Inorg. Biochem. 99, 1904–1911. [DOI] [PubMed] [Google Scholar]
- Hardman R, 2006. A toxicologic review of quantum dots: toxicity depends on physico-chemical and environmental factors. Environ. Health Perspect. 114, 165–172. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hartemann P, Hoet P, Proykova A, Fernandes T, Baun A, De Jong W, Filser J, Hensten A, Kneuer C, Maillard J-Y, 2015. Nanosilver: safety, health and environmental effects and role in antimicrobial resistance. Mater. Today 18, 122–123. [Google Scholar]
- Hornsten AM, 2001. APL-1, a Caenorhabditis elegans Protein Related to the Human Amyloid Precursor Protein, Is Essential for Viability. Boston University. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hu YO, Wang Y, Ye BP, Wang DY, 2008. Phenotypic and behavioral defects induced by iron exposure can be transferred to progeny in Caenorhabditis elegans. Biomed. Environ. Sci. 21, 467–473. [DOI] [PubMed] [Google Scholar]
- Hu C-C, Wu G-H, Lai S-F, Muthaiyan Shanmugam M, Hwu Y, Wagner OI, Yen T-J, 2018a. Toxic effects of size-tunable gold nanoparticles on Caenorhabditis elegans development and gene regulation. Sci. Rep. 8, 1–10. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hu CC, Wu GH, Hua TE, Wagner OI, Yen TJ, 2018b. Uptake of TiO(2) nanoparticles into C. elegans neurons negatively affects axonal growth and worm locomotion behavior. ACS Appl. Mater. Interfaces 10, 8485–8495. [DOI] [PubMed] [Google Scholar]
- Hu C, Hou J, Zhu Y, Lin D, 2020. Multigenerational exposure to TiO(2) nanoparticles in soil stimulates stress resistance and longevity of survived C. elegans via activating insulin/IGF-like signaling. Environ. Pollut. 263, 114376. [DOI] [PubMed] [Google Scholar]
- Huang CW, Li SW, Hsiu-Chuan Liao V, 2017. Chronic ZnO-NPs exposure at environmentally relevant concentrations results in metabolic and locomotive toxicities in Caenorhabditis elegans. Environ. Pollut. 220, 1456–1464. [DOI] [PubMed] [Google Scholar]
- Huang C-W, Li S-W, Liao VH-C, 2019. Long-term sediment exposure to ZnO nanoparticles induces oxidative stress in Caenorhabditis elegans. Environ. Sci. Nano 6, 2602–2614. [Google Scholar]
- Ijomone OM, Miah MR, Akingbade GT, Bucinca H, Aschner M, 2020. Nickel-induced developmental neurotoxicity in C. elegans includes cholinergic, dopaminergic and GABAergic degeneration, altered behaviour, and increased SKN-1 activity. Neurotox. Res. 37, 1018–1028. [DOI] [PubMed] [Google Scholar]
- Ivask A, Bondarenko O, Jepihhina N, Kahru A, 2010. Profiling of the reactive oxygen species-related ecotoxicity of CuO, ZnO, TiO2, silver and fullerene nanoparticles using a set of recombinant luminescent Escherichia coli strains: differentiating the impact of particles and solubilised metals. Anal. Bioanal. Chem. 398, 701–716. [DOI] [PubMed] [Google Scholar]
- James SA, Roberts BR, Hare DJ, de Jonge MD, Birchall IE, Jenkins NL, Cherny RA, Bush AI, McColl G, 2015. Direct in vivo imaging of ferrous iron dyshomeostasis in ageing Caenorhabditis elegans. Chem. Sci. 6, 2952–2962. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jenkins NL, James SA, Salim A, Sumardy F, Speed TP, Conrad M, Richardson DR, Bush AI, McColl G, 2020. Changes in ferrous iron and glutathione promote ferroptosis and frailty in aging Caenorhabditis elegans. Elife 9, e56580. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jiang X, Gao H, 2016. Neurotoxicity of Nanomaterials and Nanomedicine. Academic Press. [Google Scholar]
- Jiang Y, Huo S, Mizuhara T, Das R, Lee Y-W, Hou S, Moyano DF, Duncan B, Liang X-J, Rotello VM, 2015. The interplay of size and surface functionality on the cellular uptake of sub-10 nm gold nanoparticles. ACS Nano 9, 9986–9993. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jiang J, Pi J, Cai J, 2018. The advancing of zinc oxide nanoparticles for biomedical applications. Bioinorg. Chem. Appl. 2018, 1062562. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jin S, Ma X, Ma H, Zheng K, Liu J, Hou S, Meng J, Wang PC, Wu X, Liang X-J, 2013. Surface chemistry-mediated penetration and gold nanorod thermotherapy in multicellular tumor spheroids. Nanoscale 5, 143–146. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jomova K, Makova M, Alomar SY, Alwasel SH, Nepovimova E, Kuca K, Rhodes CJ, Valko M, 2022. Essential metals in health and disease. Chem. Biol. Interact. 367, 110173. [DOI] [PubMed] [Google Scholar]
- Juzenas P, Chen W, Sun Y-P, Coelho MAN, Generalov R, Generalova N, Christensen IL, 2008. Quantum dots and nanoparticles for photodynamic and radiation therapies of cancer. Adv. Drug Deliv. Rev. 60, 1600–1614. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kahison MA, Dixon SJ, 2022. Copper-Induced Cell Death Excess Copper Causes Mitochondrial Protein Aggregation and Triggers a Distinct Form of Cell Death. Amer Assoc Advancement Science, NW, WASHINGTON, DC, USA, pp. 1231–1232. [Google Scholar]
- Karlsson HL, Gustafsson J, Cronholm P, Möller L, 2009. Size-dependent toxicity of metal oxide particles—a comparison between nano-and micrometer size. Toxicol. Lett. 188, 112–118. [DOI] [PubMed] [Google Scholar]
- Kaur SJ, McKeown SR, Rashid S, 2016. Mutant SOD1 mediated pathogenesis of Amyotrophic Lateral Sclerosis. Gene 577, 109–118. [DOI] [PubMed] [Google Scholar]
- Ke T, Bornhorst J, Schwerdtle T, Santamaría A, Soare FAA, Rocha JBT, Farina M, Bowman AB, Aschner M, 2020a. Therapeutic efficacy of the N,N’ bis-(2-mercaptoethyl) isophthalamide chelator for methylmercury intoxication in Caenorhabditis elegans. Neurotox. Res. 38, 133–144. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ke T, Tsatsakis A, Santamaría A, Antunes Soare FA, Tinkov AA, Docea AO, Skalny A, Bowman AB, Aschner M, 2020b. Chronic exposure to methylmercury induces puncta formation in cephalic dopaminergic neurons in Caenorhabditis elegans. Neurotoxicology 77, 105–113. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ke T, Prince LM, Bowman AB, Aschner M, 2021. Latent alterations in swimming behavior by developmental methylmercury exposure are modulated by the homolog of tyrosine hydroxylase in Caenorhabditis elegans. Neurotoxicol. Teratol. 85, 106963. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Keller AA, McFerran S, Lazareva A, Suh S, 2013. Global life cycle releases of engineered nanomaterials. J. Nanopart. Res. 15, 1–17. [Google Scholar]
- Klang IM, Schilling B, Sorensen DJ, Sahu AK, Kapahi P, Andersen JK, Swoboda P, Killilea DW, Gibson BW, Lithgow GJ, 2014. Iron promotes protein insolubility and aging in C. elegans. Aging (Albany NY) 6, 975–991. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Klein GL, 2019. Aluminum toxicity to bone: a multisystem effect? Osteoporos Sarcopenia 5, 2–5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kong L, Gao X, Zhu J, Zhang T, Xue Y, Tang M, 2017. Reproductive toxicity induced by nickel nanoparticles in Caenorhabditis elegans. Environ. Toxicol. 32, 1530–1538. [DOI] [PubMed] [Google Scholar]
- Kumar V, Gill KD, 2009. Aluminium neurotoxicity: neurobehavioural and oxidative aspects. Arch. Toxicol. 83, 965–978. [DOI] [PubMed] [Google Scholar]
- Kumar S, Singh RK, Aman AK, Kumar J, Kar M, 2017. Evaluation of iron oxide nanoparticles (nps) on aging and age related metabolism and physiological changes in C. elegans. Int. J. Pharm. Sci. Res. 8, 3813–3816. [Google Scholar]
- Kumar A, Kumar A, Cabral-Pinto MMS, Chaturvedi AK, Shabnam AA, Subrahmanyam G, Mondal R, Gupta DK, Malyan SK, Kumar SS, 2020. Lead toxicity: health hazards, influence on food chain, and sustainable remediation approaches. Int. J. Environ. Res. Public Health 17, 2179. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lai C-H, Chou C-C, Chuang H-C, Lin G-J, Pan C-H, Chen W-L, 2020. Receptor for advanced glycation end products in relation to exposure to metal fumes and polycyclic aromatic hydrocarbon in shipyard welders. Ecotoxicol. Environ. Saf. 202, 110920. [DOI] [PubMed] [Google Scholar]
- Lawes M, Pinkas A, Frohlich BA, Iroegbu JD, Ijomone OM, Aschner M, 2020. Metal-induced neurotoxicity in a RAGE-expressing C. elegans model. Neurotoxicology 80, 71–75. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Li Y, Wang W, Wu Q, Li Y, Tang M, Ye B, Wang D, 2012a. Molecular control of TiO₂-NPs toxicity formation at predicted environmental relevant concentrations by Mn-SODs proteins. PloS One 7, e44688. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Li Y, Yu S, Wu Q, Tang M, Pu Y, Wang D, 2012b. Chronic Al2O3-nanoparticle exposure causes neurotoxic effects on locomotion behaviors by inducing severe ROS production and disruption of ROS defense mechanisms in nematode Caenorhabditis elegans. J. Hazard. Mater. 219, 221–230. [DOI] [PubMed] [Google Scholar]
- Li WH, Shi YC, Tseng IL, Liao VH, 2013a. Protective efficacy of selenite against lead-induced neurotoxicity in Caenorhabditis elegans. PloS One 8, e62387. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Li Y, Yu S, Wu Q, Tang M, Wang D, 2013b. Transmissions of serotonin, dopamine, and glutamate are required for the formation of neurotoxicity from Al2O3-NPs in nematode Caenorhabditis elegans. Nanotoxicology 7, 1004–1013. [DOI] [PubMed] [Google Scholar]
- Li W, Wang D, Wang D, 2018. Regulation of the response of Caenorhabditis elegans to simulated microgravity by p38 mitogen-activated protein kinase signaling. Sci. Rep. 8, 1–10. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Li D, Ji J, Yuan Y, Wang D, 2020. Toxicity comparison of nanopolystyrene with three metal oxide nanoparticles in nematode Caenorhabditis elegans. Chemosphere 245, 125625. [DOI] [PubMed] [Google Scholar]
- Liang S, Duan J, Hu H, Zhang J, Gao S, Jing H, Li G, Sun Z, 2020a. Comprehensive analysis of SiNPs on the genome-wide transcriptional changes in Caenorhabditis elegans. Int. J. Nanomedicine 15, 5227–5237. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Liang X, Wang Y, Cheng J, Ji Q, Wang Y, Wu T, Tang M, 2020b. Mesoporous silica nanoparticles at predicted environmentally relevant concentrations cause impairments in GABAergic motor neurons of nematode Caenorhabditis elegans. Chem. Res. Toxicol. 33, 1665–1676. [DOI] [PubMed] [Google Scholar]
- Liang X, Wang X, Cheng J, Zhang X, Wu T, 2022. Ag(2)Se quantum dots damage the nervous system of nematode Caenorhabditis elegans. Bull. Environ. Contam. Toxicol. 109, 279–285. [DOI] [PubMed] [Google Scholar]
- Liao VH, Liu JT, Li WH, Yu CW, Hsieh YC, 2010. Caenorhabditis elegans bicarbonate transporter ABTS-1 is involved in arsenite toxicity and cholinergic signaling. Chem. Res. Toxicol. 23, 926–932. [DOI] [PubMed] [Google Scholar]
- Liberati NT, Fitzgerald KA, Kim DH, Feinbaum R, Golenbock DT, Ausubel FM, 2004. Requirement for a conserved Toll/interleukin-1 resistance domain protein in the Caenorhabditis elegans immune response. Proc. Natl. Acad. Sci. U.S.A. 101, 6593–6598. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Liu J, Lu L, Xu S, Wang L, 2015. One-pot synthesis of gold nanoclusters with bright red fluorescence and good biorecognition abilities for visualization fluorescence enhancement detection of E. coli. Talanta 134, 54–59. [DOI] [PubMed] [Google Scholar]
- Liu L, He S, Tang M, Zhang M, Wang C, Wang Z, Sun F, Yan Y, Li H, Lin K, 2022. Pseudo toxicity abatement effect of norfloxacin and copper combined exposure on Caenorhabditis elegans. Chemosphere 287, 132019. [DOI] [PubMed] [Google Scholar]
- Lu C, Svoboda KR, Lenz KA, Pattison C, Ma H, 2018. Toxicity interactions between manganese (Mn) and lead (Pb) or cadmium (Cd) in a model organism the nematode C. elegans. Environ. Sci. Pollut. Res. Int. 25, 15378–15389. [DOI] [PubMed] [Google Scholar]
- Ma X, Wu Y, Jin S, Tian Y, Zhang X, Zhao Y, Yu L, Liang X-J, 2011. Gold nanoparticles induce autophagosome accumulation through size-dependent nanoparticle uptake and lysosome impairment. ACS Nano 5, 8629–8639. [DOI] [PubMed] [Google Scholar]
- Mabrouk M, Das DB, Salem ZA, Beherei HH, 2021. Nanomaterials for biomedical applications: production, characterisations, recent trends and difficulties. Molecules 26, 1077. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mao Z, Zheng Y-L, Zhang Y-Q, 2010. Behavioral impairment and oxidative damage induced by chronic application of nonylphenol. Int. J. Mol. Sci. 12, 114–127. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Maret W, 2016. The metals in the biological periodic system of the elements: concepts and conjectures. Int. J. Mol. Sci. 17, 66. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Martinez-Finley EJ, Chakraborty S, Slaughter JC, Aschner M, 2013. Early-life exposure to methylmercury in wildtype and pdr-1/parkin knockout C. elegans. Neurochem. Res. 38, 1543–1552. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Martins AC, Virgolini MB, Tinkov AA, Skalny AV, Tirumala RP, Farina M, Santamaria A, Lu R, Aschner M, 2022. Iron overload and neurodegenerative diseases: what can we learn from Caenorhabditis elegans? Toxicol. Res. Appl. 6, 23978473221091852. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mashock MJ, Zanon T, Kappell AD, Petrella LN, Andersen EC, Hristova KR, 2016. Copper oxide nanoparticles impact several toxicological endpoints and cause neurodegeneration in Caenorhabditis elegans. PloS One 11, e0167613. [DOI] [PMC free article] [PubMed] [Google Scholar]
- McElwee MK, Freedman JH, 2011. Comparative toxicology of mercurials in Caenorhabditis elegans. Environ. Toxicol. Chem. 30, 2135–2141. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mead MN, 2010. Cadmium confusion: do consumers need protection? Environ. Health Perspect. 118, a528–a534. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mo J, Sun N, Yang B, Li S, Wang L, 2020. The nervous system is the major target for gold nanoparticles: evidence from RNA sequencing data of C. elegans. bioRxiv, 699785. [Google Scholar]
- Moyson S, Vissenberg K, Fransen E, Blust R, Husson SJ, 2018. Mixture effects of copper, cadmium, and zinc on mortality and behavior of Caenorhabditis elegans. Environ. Toxicol. Chem. 37, 145–159. [DOI] [PubMed] [Google Scholar]
- M’rad I, Jeljeli M, Rihane N, Hilber P, Sakly M, Amara S, 2018. Aluminium oxide nanoparticles compromise spatial learning and memory performance in rats. EXCLI J. 17, 200. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mushtaq G, Khan JA, Joseph E, Kamal MA, 2015. Nanoparticles, neurotoxicity and neurodegenerative diseases. Curr. Drug Metab. 16, 676–684. [DOI] [PubMed] [Google Scholar]
- Nguyen JP, Shipley FB, Linder AN, Plummer GS, Liu M, Setru SU, Shaevitz JW, Leifer AM, 2016. Whole-brain calcium imaging with cellular resolution in freely behaving Caenorhabditis elegans. Proc. Natl. Acad. Sci. U.S.A. 113, E1074–E1081. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Nicolai MM, Pirritano M, Gasparoni G, Aschner M, Simon M, Bornhorst J, 2022. Manganese-induced toxicity in C. elegans: what can we learn from the transcriptome? Int. J. Mol. Sci. 23, 10748. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Nohynek GJ, Antignac E, Re T, Toutain H, 2010. Safety assessment of personal care products/cosmetics and their ingredients. Toxicol. Appl. Pharmacol. 243, 239–259. [DOI] [PubMed] [Google Scholar]
- Nowack B, 2017. Evaluation of environmental exposure models for engineered nanomaterials in a regulatory context. NanoImpact 8, 38–47. [Google Scholar]
- Nowack B, Krug HF, Height M, 2011. 120 years of nanosilver history: implications for policy makers. Environ. Sci. Technol. 45, 1177–1183. [DOI] [PubMed] [Google Scholar]
- Ortega R, Carmona A, 2022. Neurotoxicity of Environmental Metal Toxicants. MDPI, p. 382. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Page KE, White KN, McCrohan CR, Killilea DW, Lithgow GJ, 2012. Aluminium exposure disrupts elemental homeostasis in Caenorhabditis elegans. Metallomics 4, 512–522. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Park B, Donaldson K, Duffin R, Tran L, Kelly F, Mudway I, Morin JP, Guest R, Jenkinson P, Samaras Z, Giannouli M, Kouridis H, Martin P, 2008. Hazard and risk assessment of a nanoparticulate cerium oxide-based diesel fuel additive - a case study. Inhal. Toxicol. 20, 547–566. [DOI] [PubMed] [Google Scholar]
- Patel D, Xu C, Nagarajan S, Liu Z, Hemphill WO, Shi R, Uversky VN, Caldwell GA, Caldwell KA, Witt SN, 2018. Alpha-synuclein inhibits Snx3-retromer-mediated retrograde recycling of iron transporters in S. cerevisiae and C. elegans models of Parkinson’s disease. Hum. Mol. Genet. 27, 1514–1532. [DOI] [PubMed] [Google Scholar]
- Peralta-Videa JR, Zhao L, Lopez-Moreno ML, de la Rosa G, Hong J, Gardea-Torresdey JL, 2011. Nanomaterials and the environment: a review for the biennium 2008–2010. J. Hazard. Mater. 186, 1–15. [DOI] [PubMed] [Google Scholar]
- Pereira FSO, Barbosa FAR, Canto RFS, Lucchese C, Pinton S, Braga AL, Azeredo JB, Quines CB, Ávila DS, 2022. Dihydropyrimidinone-derived selenoesters efficacy and safety in an in vivo model of Aβ aggregation. Neurotoxicology 88, 14–24. [DOI] [PubMed] [Google Scholar]
- Peres TV, Arantes LP, Miah MR, Bornhorst J, Schwerdtle T, Bowman AB, Leal RB, Aschner M, 2018. Role of Caenorhabditis elegans AKT-1/2 and SGK-1 in manganese toxicity. Neurotox. Res. 34, 584–596. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Peres TV, Horning KJ, Bornhorst J, Schwerdtle T, Bowman AB, Aschner M, 2019. Small molecule modifiers of in vitro manganese transport alter toxicity in vivo. Biol. Trace Elem. Res. 188, 127–134. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Peterson RT, Nass R, Boyd WA, Freedman JH, Dong K, Narahashi T, 2008. Use of non-mammalian alternative models for neurotoxicological study. Neurotoxicology 29, 546–555. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pihl RO, Parkes M, 1977. Hair element content in learning disabled children. Science 198, 204–206. [DOI] [PubMed] [Google Scholar]
- Pluskota A, Horzowski E, Bossinger O, von Mikecz A, 2009. In Caenorhabditis elegans nanoparticle-bio-interactions become transparent: silica-nanoparticles induce reproductive senescence. PloS One 4, e6622. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Prasad R, Rattan G, 2010. Preparation methods and applications of CuO-CeO2 catalysts: a short review. Bull. Chem. React. Eng. Catal. 5, 7. [Google Scholar]
- Pulit-Prociak J, Banach M, 2016. Silver nanoparticles—a material of the future…? Open Chem. 14, 76–91. [Google Scholar]
- Queirós L., Pereira J., Gonçalves F., Pacheco M, Aschner M., Pereira P., 2019. Caenorhabditis elegans as a tool for environmental risk assessment: emerging and promising applications for a “nobelized worm”. Crit. Rev. Toxicol. 49, 411–429. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Raizen D, Song B. m., Trojanowski N, You Y-J, 2018. Methods for measuring pharyngeal behaviors. In: The C. elegans Research Community (Ed.), WormBook: The Online Review of C. elegans Biology [Internet]. Wormbook. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Raj V, Nair A, Thekkuveettil A, 2021. Manganese exposure during early larval stages of C. elegans causes learning disability in the adult stage. Biochem. Biophys. Res. Commun. 568, 89–94. [DOI] [PubMed] [Google Scholar]
- Rogers S, Rice KM, Manne ND, Shokuhfar T, He K, Selvaraj V, Blough ER, 2015. Cerium oxide nanoparticle aggregates affect stress response and function in Caenorhabditis elegans. SAGE Open Med. 3, 2050312115575387. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Rzigalinski BA, Strobl JS, 2009. Cadmium-containing nanoparticles: perspectives on pharmacology and toxicology of quantum dots. Toxicol. Appl. Pharmacol. 238, 280–288. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Scharf A, Piechulek A, von Mikecz A, 2013. Effect of nanoparticles on the biochemical and behavioral aging phenotype of the nematode Caenorhabditis elegans. ACS Nano 7, 10695–10703. [DOI] [PubMed] [Google Scholar]
- Scharf A, Gührs K-H, von Mikecz A, 2016. Anti-amyloid compounds protect from silica nanoparticle-induced neurotoxicity in the nematode C. elegans. Nanotoxicology 10, 426–435. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Schetinger MRC, Peres TV, Arantes LP, Carvalho F, Dressler V, Heidrich G, Bowman AB, Aschner M, 2019. Combined exposure to methylmercury and manganese during L1 larval stage causes motor dysfunction, cholinergic and monoaminergic up-regulation and oxidative stress in L4 Caenorhabditis elegans. Toxicology 411, 154–162. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Schilling K, Bradford B, Castelli D, Dufour E, Nash JF, Pape W, Schulte S, Tooley I, van den Bosch J, Schellauf F, 2010. Human safety review of “nano” titanium dioxide and zinc oxide. Photochem. Photobiol. Sci. 9, 495–509. [DOI] [PubMed] [Google Scholar]
- Seabra AB, Durán N, 2015. Nanotoxicology of Metal Oxide Nanoparticles. Metals 5, 934–975. [Google Scholar]
- Sedensky MM, Morgan PG, 2018. Using Caenorhabditis elegans to study neurotoxicity. In: Handbook of Developmental Neurotoxicology Elsevier, pp. 153–160. [Google Scholar]
- Selkoe DJ, 2001. Alzheimer’s disease: genes, proteins, and therapy. Physiol. Rev. 81, 741–766. [DOI] [PubMed] [Google Scholar]
- Shahbazi MA, Herranz B, Santos HA, 2012. Nanostructured porous Si-based nanoparticles for targeted drug delivery. Biomatter 2, 296–312. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Shang L, Gardner DP, Xu W, Cannone JJ, Miranker DP, Ozer S, Gutell RR, 2013. Two accurate sequence, structure, and phylogenetic template-based RNA alignment systems. BMC Syst. Biol. 7, 1–15. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sidoryk-Wegrzynowicz M, Aschner M, 2013. Role of astrocytes in manganese mediated neurotoxicity. BMC Pharmacol. Toxicol. 14, 23. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Singh N, Jenkins GJ, Asadi R, Doak SH, 2010. Potential toxicity of superparamagnetic iron oxide nanoparticles (SPION). Nano Rev. 1, 5358. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sinis SI, Gourgoulianis KI, Hatzoglou C, Zarogiannis SG, 2019. Mechanisms of engineered nanoparticle induced neurotoxicity in Caenorhabditis elegans. Environ. Toxicol. Pharmacol. 67, 29–34. [DOI] [PubMed] [Google Scholar]
- Smith AM, Duan H, Mohs AM, Nie S, 2008. Bioconjugated quantum dots for in vivo molecular and cellular imaging. Adv. Drug Deliv. Rev. 60, 1226–1240. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Soares FA, Fagundez DA, Avila DS, 2017. Neurodegeneration induced by metals in Caenorhabditis elegans. Adv. Neurobiol. 18, 355–383. [DOI] [PubMed] [Google Scholar]
- Song S, Guo Y, Zhang X, Zhang X, Zhang J, Ma E, 2014. Changes to cuticle surface ultrastructure and some biological functions in the nematode Caenorhabditis elegans exposed to excessive copper. Arch. Environ. Contam. Toxicol. 66, 390–399. [DOI] [PubMed] [Google Scholar]
- Sørensen SN, Baun A, 2015. Controlling silver nanoparticle exposure in algal toxicity testing—a matter of timing. Nanotoxicology 9, 201–209. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sulston J, 1983. Neuronal Cell Lineages in the Nematode Caenorhabditis elegans, Cold Spring Harbor symposia on quantitative biology. Cold Spring Harbor Laboratory Press, pp. 443–452. [DOI] [PubMed] [Google Scholar]
- Syafiuddin A, Salim MR, Beng Hong Kueh A, Hadibarata T, Nur H, 2017. A review of silver nanoparticles: research trends, global consumption, synthesis, properties, and future challenges. J. Chin. Chem. Soc. 64, 732–756. [Google Scholar]
- Tang B, Tong P, Xue KS, Williams PL, Wang JS, Tang L, 2019. High-throughput assessment of toxic effects of metal mixtures of cadmium(Cd), lead(Pb), and manganese(Mn) in nematode Caenorhabditis elegans. Chemosphere 234, 232–241. [DOI] [PubMed] [Google Scholar]
- Tao C, Zhu Y, Li X, Hanagata N, 2014. Magnetic mesoporous silica nanoparticles for CpG delivery to enhance cytokine induction via toll-like receptor 9. RSC Adv. 4, 45823–45830. [Google Scholar]
- Tayel AA, El-Tras WF, Moussa S, El-Baz AF, Mahrous H, Salem MF, Brimer L, 2011. Antibacterial action of zinc oxide nanoparticles against foodborne pathogens. J. Food Saf. 31, 211–218. [Google Scholar]
- Tchounwou PB, Yedjou CG, Patlolla AK, Sutton DJ, 2012. Heavy metal toxicity and the environment. Exp. Suppl. 101, 133–164. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Teleanu DM, Chircov C, Grumezescu AM, Teleanu RI, 2019. Neurotoxicity of nanomaterials: an up-to-date overview. Nanomaterials (Basel) 9, 96. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tiernan CT, Edwin EA, Goudreau JL, Atchison WD, Lookingland KJ, 2013. The role of de novo catecholamine synthesis in mediating methylmercury-induced vesicular dopamine release from rat pheochromocytoma (PC12) cells. Toxicol. Sci. 133, 125–132. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tietz T, Lenzner A, Kolbaum AE, Zellmer S, Riebeling C, Gurtler R, Jung C, Kappenstein O, Tentschert J, Giulbudagian M, Merkel S, Pirow R, Lindtner O, Tralau T, Schafer B, Laux P, Greiner M, Lampen A, Luch A, Wittkowski R, Hensel A, 2019. Aggregated aluminium exposure: risk assessment for the general population. Arch. Toxicol. 93, 3503–3521. [DOI] [PubMed] [Google Scholar]
- Tortora GJ, Derrickson BH, 2018. Principles of Anatomy and Physiology. John Wiley & Sons. [Google Scholar]
- Tsyusko OV, Unrine JM, Spurgeon D, Blalock E, Starnes D, Tseng M, Joice G, Bertsch PM, 2012. Toxicogenomic responses of the model organism Caenorhabditis elegans to gold nanoparticles. Environ. Sci. Technol. 46, 4115–4124. [DOI] [PubMed] [Google Scholar]
- Umair M, Javed I, Rehman M, Madni A, Javeed A, Ghafoor A, Ashraf M, 2016. Nanotoxicity of inert materials: the case of gold, silver and iron. J. Pharm. Pharm. Sci. 19, 161–180. [DOI] [PubMed] [Google Scholar]
- USDA, 2019. FoodData Central, Unites States Deparment of Agriculture. [Google Scholar]
- Valdiglesias V, Fernández-Bertólez N, Kilic G., Costa C., Costa S., Fraga S, Bessa MJ., Pásaro E, Teixeira JP, Laffon B, 2016. Are iron oxide nanoparticles safe? Current knowledge and future perspectives. J. Trace Elem. Med. Biol. 38, 53–63. [DOI] [PubMed] [Google Scholar]
- Vanduyn N, Settivari R, Wong G, Nass R, 2010. SKN-1/Nrf2 inhibits dopamine neuron degeneration in a Caenorhabditis elegans model of methylmercury toxicity. Toxicol. Sci. 118, 613–624. [DOI] [PMC free article] [PubMed] [Google Scholar]
- VanDuyn N, Settivari R, LeVora J, Zhou S, Unrine J, Nass R, 2013. The metal transporter SMF-3/DMT-1 mediates aluminum-induced dopamine neuron degeneration. J. Neurochem. 124, 147–157. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Vellingiri B, Suriyanarayanan A, Selvaraj P, Abraham KS, Pasha MY, Winster H, Gopalakrishnan AV, Singaravelu G, Reddy JK, Ayyadurai N, 2022. Role of heavy metals (copper (Cu), arsenic (As), cadmium (Cd), iron (Fe) and lithium (Li)) induced neurotoxicity. Chemosphere 301, 134625. [DOI] [PubMed] [Google Scholar]
- Viau C, Haçariz O., Karimian F., Xia J., 2020. Comprehensive phenotyping and transcriptome profiling to study nanotoxicity in C. elegans. PeerJ 8, e8684. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Waalkes MP, 2000. Cadmium carcinogenesis in review. J. Inorg. Biochem. 79, 241–244. [DOI] [PubMed] [Google Scholar]
- Wang B, Du Y, 2013. Cadmium and its neurotoxic effects. Oxid. Med. Cell. Longev. 2013, 898034. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang D, Xing X, 2008. Assessment of locomotion behavioral defects induced by acute toxicity from heavy metal exposure in nematode Caenorhabditis elegans. J. Environ. Sci. (China) 20, 1132–1137. [DOI] [PubMed] [Google Scholar]
- Wang DY, Yang YC, Wang Y, 2009a. Aluminium toxicosis causing transferable defects from exposed animals to their progeny in Caenorhabditis elegans. Zhonghua Yu Fang Yi Xue Za Zhi 43, 45–51. [PubMed] [Google Scholar]
- Wang H, Wick RL, Xing B, 2009b. Toxicity of nanoparticulate and bulk ZnO, Al2O3 and TiO2 to the nematode Caenorhabditis elegans. Environ. Pollut. 157, 1171–1177. [DOI] [PubMed] [Google Scholar]
- Wang Y, Chen J, Irudayaraj J, 2011. Nuclear targeting dynamics of gold nanoclusters for enhanced therapy of HER2+ breast cancer. ACS Nano 5, 9718–9725. [DOI] [PubMed] [Google Scholar]
- Wang S, Chu Z, Zhang K, Miao G, 2018. Cadmium-induced serotonergic neuron and reproduction damages conferred lethality in the nematode Caenorhabditis elegans. Chemosphere 213, 11–18. [DOI] [PubMed] [Google Scholar]
- White J, Southgate E, Thomson JN, Brenner S, 1986. The structure of the nervous system of the nematode C. elegans. Philos. Trans. R. Soc. Lond. B Biol. Sci. 314, 1–340. [DOI] [PubMed] [Google Scholar]
- Win-Shwe T-T, Fujimaki H, 2011. Nanoparticles and neurotoxicity. Int. J. Mol. Sci. 12, 6267–6280. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Witkowska D, Slowik J, Chilicka K, 2021. Heavy metals and human health: possible exposure pathways and the competition for protein binding sites. Molecules 26, 6060. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wu S, Lu J, Rui Q, Yu S, Cai T, Wang D, 2011. Aluminum nanoparticle exposure in L1 larvae results in more severe lethality toxicity than in L4 larvae or young adults by strengthening the formation of stress response and intestinal lipofuscin accumulation in nematodes. Environ. Toxicol. Pharmacol. 31, 179–188. [DOI] [PubMed] [Google Scholar]
- Wu Q, Li Y, Tang M, Wang D, 2012a. Evaluation of environmental safety concentrations of DMSA coated Fe2O3-NPs using different assay systems in nematode Caenorhabditis elegans. PLoS ONE 7, e43729. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wu Q, Liu P, Li Y, Du M, Xing X, Wang D, 2012b. Inhibition of ROS elevation and damage to mitochondrial function prevents lead-induced neurotoxic effects on structures and functions of AFD neurons in Caenorhabditis elegans. J. Environ. Sci. (China) 24, 733–742. [DOI] [PubMed] [Google Scholar]
- Wu Q, Wang W, Li Y, Li Y, Ye B, Tang M, Wang D, 2012c. Small sizes of TiO2-NPs exhibit adverse effects at predicted environmental relevant concentrations on nematodes in a modified chronic toxicity assay system. J. Hazard. Mater. 243, 161–168. [DOI] [PubMed] [Google Scholar]
- Wu Q, Nouara A, Li Y, Zhang M, Wang W, Tang M, Ye B, Ding J, Wang D, 2013. Comparison of toxicities from three metal oxide nanoparticles at environmental relevant concentrations in nematode Caenorhabditis elegans. Chemosphere 90, 1123–1131. [DOI] [PubMed] [Google Scholar]
- Wu Q, Zhao Y, Li Y, Wang D, 2014. Susceptible genes regulate the adverse effects of TiO2-NPs at predicted environmental relevant concentrations on nematode Caenorhabditis elegans. Nanomedicine 10, 1263–1271. [DOI] [PubMed] [Google Scholar]
- World Health Organization, 2022. Chapter 5. Chemicals. [Google Scholar]
- Wu T, He K, Zhan Q, Ang S, Ying J, Zhang S, Zhang T, Xue Y, Tang M, 2015. MPA-capped CdTe quantum dots exposure causes neurotoxic effects in nematode Caenorhabditis elegans by affecting the transporters and receptors of glutamate, serotonin and dopamine at the genetic level, or by increasing ROS, or both. Nanoscale 7, 20460–20473. [DOI] [PubMed] [Google Scholar]
- Wu T, Xu H, Liang X, Tang M, 2019. Caenorhabditis elegans as a complete model organism for biosafety assessments of nanoparticles. Chemosphere 221, 708–726. [DOI] [PubMed] [Google Scholar]
- Wu J, Yang JJ, Cao Y, Li H, Zhao H, Yang S, Li K, 2020. Iron overload contributes to general anaesthesia-induced neurotoxicity and cognitive deficits. J. Neuroinflammation 17, 110. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Xabier MG, Howe RC, Frommel D, 1992. Infection and elimination of Mycobacterium leprae in SCID C.B.−17 mice (severe combined immunodeficiency). C. R. Acad. Sci. III 314, 99–103. [PubMed] [Google Scholar]
- Xing X, Du M, Zhang Y, Wang D, 2009a. Adverse effects of metal exposure on chemotaxis towards water-soluble attractants regulated mainly by ASE sensory neuron in nematode Caenorhabditis elegans. J. Environ. Sci. (China) 21, 1684–1694. [DOI] [PubMed] [Google Scholar]
- Xing XJ, Rui Q, Du M, Wang DY, 2009b. Exposure to lead and mercury in young larvae induces more severe deficits in neuronal survival and synaptic function than in adult nematodes. Arch. Environ. Contam. Toxicol. 56, 732–741. [DOI] [PubMed] [Google Scholar]
- Yang YF, Cheng YH, Liao CM, 2016. In situ remediation-released zero-valent iron nanoparticles impair soil ecosystems health: a C. elegans biomarker-based risk assessment. J. Hazard. Mater. 317, 210–220. [DOI] [PubMed] [Google Scholar]
- Yang Y, Xu G, Xu S, Chen S, Xu A, Wu L, 2018. Effect of ionic strength on bioaccumulation and toxicity of silver nanoparticles in Caenorhabditis elegans. Ecotoxicol. Environ. Saf. 165, 291–298. [DOI] [PubMed] [Google Scholar]
- Ye H, Ye B, Wang D, 2008. Trace administration of vitamin E can retrieve and prevent UV-irradiation- and metal exposure-induced memory deficits in nematode Caenorhabditis elegans. Neurobiol. Learn. Mem. 90, 10–18. [DOI] [PubMed] [Google Scholar]
- Younis A, Chu D, Li S, 2016. Cerium oxide nanostructures and their applications. Funct. Nanomater. 3, 53–68. [Google Scholar]
- Yu S, Rui Q, Cai T, Wu Q, Li Y, Wang D, 2011. Close association of intestinal autofluorescence with the formation of severe oxidative damage in intestine of nematodes chronically exposed to Al2O3-nanoparticle. Environ. Toxicol. Pharmacol. 32, 233–241. [DOI] [PubMed] [Google Scholar]
- Zambrano N, Bimonte M, Arbucci S, Gianni D, Russo T, Bazzicalupo P, 2002. feh-1 and apl-1, the Caenorhabditis elegans orthologues of mammalian Fe65 and β-amyloid precursor protein genes, are involved in the same pathway that controls nematode pharyngeal pumping. J. Cell Sci. 115, 1411–1422. [DOI] [PubMed] [Google Scholar]
- Zhang P, Kong J, 2015. Doxorubicin-tethered fluorescent silica nanoparticles for pH-responsive anticancer drug delivery. Talanta 134, 501–507. [DOI] [PubMed] [Google Scholar]
- Zhang H, He X, Zhang Z, Zhang P, Li Y, Ma Y, Kuang Y, Zhao Y, Chai Z, 2011. Nano-CeO2 exhibits adverse effects at environmental relevant concentrations. Environ. Sci. Technol. 45, 3725–3730. [DOI] [PubMed] [Google Scholar]
- Zhang XD, Luo Z, Chen J, Shen X, Song S, Sun Y, Fan S, Fan F, Leong DT, Xie J, 2014. Ultrasmall Au10–12 (SG) 10–12 nanomolecules for high tumor specificity and cancer radiotherapy. Adv. Mater. 26, 4565–4568. [DOI] [PubMed] [Google Scholar]
- Zhang X, Zhong HQ, Chu ZW, Zuo X, Wang L, Ren XL, Ma H, Du RY, Ju JJ, Ye XL, Huang CP, Zhu JH, Wu HM, 2020. Arsenic induces transgenerational behavior disorders in Caenorhabditis elegans and its underlying mechanisms. Chemosphere 252, 126510. [DOI] [PubMed] [Google Scholar]
- Zhang S, Chu Q, Zhang Z, Xu Y, Mao X, Zhang M, 2021a. Responses of Caenorhabditis elegans to various surface modifications of alumina nanoparticles. Environ. Pollut. 271, 116335. [DOI] [PubMed] [Google Scholar]
- Zhang W, Li W, Li J, Chang X, Niu S, Wu T, Kong L, Zhang T, Tang M, Xue Y, 2021b. Neurobehavior and neuron damage following prolonged exposure of silver nanoparticles with/without polyvinylpyrrolidone coating in Caenorhabditis elegans. J. Appl. Toxicol. 41, 2055–2067. [DOI] [PubMed] [Google Scholar]
- Zhang Y, Zhao C, Zhang H, Lu Q, Zhou J, Liu R, Wang S, Pu Y, Yin L, 2021c. Trans-generational effects of copper on nerve damage in Caenorhabditis elegans. Chemosphere 284, 131324. [DOI] [PubMed] [Google Scholar]
- Zhao Y, Wu Q, Tang M, Wang D, 2014. The in vivo underlying mechanism for recovery response formation in nano-titanium dioxide exposed Caenorhabditis elegans after transfer to the normal condition. Nanomedicine 10, 89–98. [DOI] [PubMed] [Google Scholar]
- Zhao Y, Wang X, Wu Q, Li Y, Wang D, 2015. Translocation and neurotoxicity of CdTe quantum dots in RMEs motor neurons in nematode Caenorhabditis elegans. J. Hazard. Mater. 283, 480–489. [DOI] [PubMed] [Google Scholar]
- Zhao Y, Zhi L, Wu Q, Yu Y, Sun Q, Wang D, 2016. p38 MAPK-SKN-1/Nrf signaling cascade is required for intestinal barrier against graphene oxide toxicity in Caenorhabditis elegans. Nanotoxicology 10, 1469–1479. [DOI] [PubMed] [Google Scholar]
- Zheng H, Koo EH, 2011. Biology and pathophysiology of the amyloid precursor protein. Mol. Neurodegener. 6, 1–16. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zheng W, Aschner M, Ghersi-Egea JF, 2003. Brain barrier systems: a new frontier in metal neurotoxicological research. Toxicol. Appl. Pharmacol. 192, 1–11. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zheng F, Chen P, Li H, Aschner M, 2020. Drp-1-dependent mitochondrial fragmentation contributes to cobalt chloride-induced toxicity in Caenorhabditis elegans. Toxicol. Sci. 177, 158–167. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhou RP, Chen Y, Wei X, Yu B, Xiong ZG, Lu C, Hu W, 2020. Novel insights into ferroptosis: implications for age-related diseases. Theranostics 10, 11976–11997. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zoroddu MA, Aaseth J, Crisponi G, Medici S, Peana M, Nurchi VM, 2019. The essential metals for humans: a brief overview. J. Inorg. Biochem. 195, 120–129. [DOI] [PubMed] [Google Scholar]