J. Pineal Res. 2014

© 2014 John Wiley & Sons A/S. Published by John Wiley & Sons Ltd

Doi:10.1111/jpi.12132

Journal of Pineal Research

Molecular, Biological, Physiological and Clinical Aspects of Melatonin

REVIEW ARTICLE

A review of metal-catalyzed molecular damage: protection by melatonin Abstract: Metal exposure is associated with several toxic effects; herein, we review the toxicity mechanisms of cadmium, mercury, arsenic, lead, aluminum, chromium, iron, copper, nickel, cobalt, vanadium, and molybdenum as these processes relate to free radical generation. Free radicals can be generated in cells due to a wide variety of exogenous and endogenous processes, causing modifications in DNA bases, enhancing lipid peroxidation, and altering calcium and sulfhydryl homeostasis. Melatonin, an ubiquitous and pleiotropic molecule, exerts efficient protection against oxidative stress and ameliorates oxidative/nitrosative damage by a variety of mechanisms. Also, melatonin has a chelating property which may contribute in reducing metal-induced toxicity as we postulate here. The aim of this review was to highlight the protective role of melatonin in counteracting metal-induced free radical generation. Understanding the physicochemical insights of melatonin related to the free radical scavenging activity and the stimulation of antioxidative enzymes is of critical importance for the development of novel therapeutic strategies against the toxic action of these metals.

Alejandro Romero1, Eva Ramos1,  bal de Los Rıos2,3, Javier Cristo Egea2, Javier del Pino1 and Russel J. Reiter4 Departamento de Toxicologıa y Farmacologıa, Facultad de Veterinaria, Universidad Complutense de Madrid, Madrid, Spain; 2 filo Hernando y Departamento de Instituto Teo utica, Facultad de Farmacologıa y Terape noma de Madrid, Medicina, Universidad Auto n Madrid, Spain; 3Instituto de Investigacio Sanitaria, Servicio de Farmacologıa Clınica, Hospital Universitario de la Princesa, Madrid, Spain; 4Department of Cellular and Structural Biology, University of Texas Health Science, Center at San Antonio, San Antonio, TX, USA 1

Key words: chelating properties, free radicals, melatonin, metals, oxidative stress Address reprint requests to Alejandro Romero, Departamento de Toxicologıa y Farmacologıa, Facultad de Veterinaria, Universidad Complutense de Madrid, Avda, Puerta de Hierro s/n 28040-Madrid, Spain. E-mail: [email protected] Received February 17, 2014; Accepted March 11, 2014.

Introduction Metals are crucial for a wide variety of biologic processes of living systems; however, numerous studies have provided evidence that xenobiotic metals, with no physiological functions such as aluminum, cadmium, lead, mercury, and arsenic, can interact with biologic macromolecules causing oxidative damage. Although the molecular mechanisms are not completely understood, the capacity of metals to generate reactive oxygen (ROS) and nitrogen species (RNS), and, thus, to disrupt the maintenance of redox homeostasis is considered the most important event involved in metal-induced toxicity [1]. The metal-induced ROS/RNS generation is directly involved in the induction of epigenetic changes, abnormal cell signaling and uncontrolled cell growth, initiation of cellular injury and the stimulation of inflammatory processes; each of these is pivotal mechanisms that can lead to cancer development [2, 3]. Interestingly, antioxidant therapy protects against metal toxicity by trapping free radicals and maintaining redox homeostasis. In line with this, melatonin is known to be a versatile and ubiquitously functioning molecule [4–11]. The antioxidant abilities of this indoleamine derive from both its direct scavenging of free radicals [12–17] and by increasing the activity and

expression of antioxidant enzymes [18–20]. Conversely, deficiencies in melatonin production or melatonin receptor expression and decreases in melatonin levels likely contribute to numerous dysfunctions and diseases. Thus, its loss is associated with a multitude of pathophysiological changes [21–27]. Considering its diverse actions and given melatonin’s high lipophilicity and the ease with which it crosses morphophysiological barriers, its use as a combination therapy with a wide variety of currently used metals suggests interesting therapeutic perspectives. The purpose of this review is to provide a detailed overview of the current state of knowledge related to the role of metals in the generation of ROS/RNS and tissue injury and to summarize the research performed in recent years on the protective role of melatonin against metal-induced toxicity. Furthermore, we included, from a chemical point of view, a hypothetical representation on the chelating properties of melatonin. Cadmium and melatonin The heavy and nonessential divalent metal cadmium (Cd2+) is the main toxic form of Cd that induces oxidative stress. Although Cd2+ is not a Fenton metal, there are several mechanisms by which Cd2+ indirectly can generate 1

Romero et al. ROS [28, 29]. Human exposure to Cd2+ comes from a variety of sources: drinking water, food, air, and high concentrations are found in cigarettes. Therefore, Cd2+ is an ubiquitous element present in the environment and it is classified as a Class 1 human carcinogen [30]. In an interestingly recent review by Thevenod and Lee [29], the relevant signaling pathways and action mechanisms that are targeted by Cd2+ as well as involvement of ROS signals are analyzed. In this complex scenario, it should be realized that there are different mechanisms of acute and chronic Cd2+ toxicity. It has been suggested that the mechanisms of acute Cd2+ toxicity elicit a persistent rise in ROS and Ca2+ which disrupt cell function and trigger cell death [31] in both in vivo and in vitro models. Moreover, Cd2+ inhibits antioxidative enzyme activity probably by directly binding to these proteins [32–34], causes depletion of glutathione (GSH) and protein-bound sulfhydryl groups, induces lipid peroxidation, and mediates DNA damage. Conversely, chronic low Cd2+ exposure conditions are more complex, and the roles of ROS are variable depending on experimental conditions.

It is known that ROS/RNS play a pivotal role as signaling molecules in a manner similar to other second messengers. These toxic molecules are induced by physiological or external stimuli including cytokines or metals. Also ROS/ RNS can be generated through nonenzymatic processes that are not highly regulated; they can be also produced by the activity of NADPH oxidases (NOX) which act as signaling molecules [35]. Taking into account the disruption caused by Cd2+ in numerous systems, it would be expected that melatonin would be a therapeutic multipotent agent against damage induced by this heavy metal. At the cellular level, herein, we summarize the main toxic effects and signaling cascades induced by Cd2+ and the protective role of melatonin (Fig. 1). To our knowledge, the first report that documented the protective action of melatonin against Cd2+ was conducted by Kim et al. [36] who demonstrate, in an in vivo model, that melatonin (10 mg/kg b.w., i.p.) co-administered daily with cadmium chloride (CdCl2) restored the reduction in hepatic GSH levels and ameliorated histopathological changes after Cd2+ exposure. A few years

Fig. 1. Model of cadmium pathways inducing cellular stress and the protective role of melatonin. Cd induces endoplasmic reticulum (ER) stress by calcium overload through direct activation of ceramide and binding to protein G-coupled receptors. The increased [Ca2+]c induces calpain–caspase-3-mediated apoptosis. Cd activates directly or indirectly mitochondrial reactive oxygen species (ROS)/reactive nitrogen species (RNS) generation and release the pro-apoptotic regulator cytochrome c (Cyt c) promoting caspase-3-mediated apoptosis. Cd down-regulates Ube2d genes inhibiting the P53 degradation which causes P53 overload inducing apoptosis. Melatonin counteracts Cd damage by blocking caspase-3 and reducing the ER stress caused by ROS/RNS. Melatonin also activates antioxidative enzymes and the GSH/GSSG cycle.

2

Melatonin for protection against metals later, Karbownik et al. [37] showed that lipid peroxidation induced by CdCl2 (1 mg/kg b.w., i.p.) in hamster tissues was prevented differentially by melatonin. It was presumed on the basis of these studies that melatonin reduced cadmium toxicity firstly by acting as a direct scavenger of free radicals which initiated lipid peroxidation processes and secondly by stimulating the activity of antioxidative enzymes. Thus, the pretreatment with three natural antioxidants, curcumin, resveratrol, and melatonin, in all cases, protected against Cd2+-induced lipid peroxidation and ameliorated oxidative damage in Cd2+-treated mice [38]. In addition to direct scavenging and/or the indirect antioxidative properties of melatonin, in an in vivo study, this indoleamine also reduced kidney Cd2+ accumulation [39]. This fact could be due to (i) melatonin’s lipophilic character which allows it to cross cellular membranes permitting the removal of Cd2+, (ii) establishing stable complexes with Cd2+ [40] (see hypothesis postulated herein, Fig. 2), or (iii) inhibition of intestinal absorption of Cd2+. Furthermore, melatonin administration alone or in combination with selenium and vitamin E prevents oxidative stress induced by Cd2+ in the plasma [41], testes [42], liver, and kidney of rats [43]. Knowing that Cd2+ also acts as neurotoxic agent and endocrine disruptor, previous reports have shown the protective effect of melatonin (0.4 and 3 lg/mL in the drinking water) against oxidative stress induced by very low dose of CdCl2 (5 ppm in the drinking water). In these studies, Cd2+ acted on stress marker gene expression including inducible nitric oxide synthase (iNOS) and heme oxygenase-1 (HO-1) among others, both at the hypothalamic and pituitary level [44–47]. Metallothioneins (MTs) are an important family of proteins endowed with a high capacity to bind heavy metals in biologic systems, thereby protecting against oxidative injury [48]. In an in vitro study on three cell lines, AlonsoGonzalez et al. [49] revealed that melatonin increased Cd2+-induced gene expression of MT isoform MT-2A in all cell types studied. This isoform is widely expressed in humans, and its induction by melatonin aids in reducing oxidative damage induced by Cd2+. Cellular and molecular mechanisms in Cd2+-induced nephrotoxicity have been widely studied. However, further research is necessary to identify an appropriate therapeutic approach to Cd2+-induced renal injury. Melatonin has shown a substantial ability in preventing oxidative stress and tissue damage resulting from Cd2+ toxicity [50]. The same research group exposed rats in vivo to Cd2+ (5 mg/kg b.w. s.c.) for 22 days and recorded an increase in malondialdehyde (MDA) levels and a reduction in both

Fig. 2. Hypothetical g6-coordination of metals to melatonin through the p-electron density of the benzene-fused ring in a tetrahedral-like fashion.

the activity of superoxide dismutase (SOD) and GSH concentrations in the liver. In contrast, when melatonin was administered with Cd2+, MDA levels and the low enzymatic values found after Cd2+ treatment were restored to control levels [51]. Other important targets for Cd2+ are the testes and cardiovascular system. A recent report confirmed necrosis of seminiferous tubules in Cd2+-treated mice. In this model, melatonin treatment not only alleviated Cd2+-induced histopathological injury but also attenuated testicular HO-1 up-regulation and protected against apoptosis in the testes [52]. Similarly, Cd2+-exposed rats experienced a significant decrease in myocardial antioxidative enzymes; however, simultaneous administration of melatonin and alpha-lipoic acid provided cardioprotection, minimized free radical generation, and maintained the antioxidant status [53]. Based on these findings, it is likely that the administration of exogenous melatonin may be an effective means to prevent Cd2+ toxicity. Presumably, the effective dose required to suppress Cd2+ toxicity would vary according to the amount of Cd2+ to which and individual was exposed. Mercury and melatonin Mercury is a widespread environmental and industrial pollutant; consequently, it is practically impossible for humans to avoid exposure to some of the forms of mercury. It is recognized as one of the most dangerous environmental contaminants. Humans and animals are exposed to numerous chemical forms of mercury, including elemental mercury vapor (Hg), inorganic mercurous [Hg(I)], mercuric [Hg(II)] and organic mercuric compounds [28, 54], all of which induce severe alterations in tissues of both animals and man. Also, it functions as an endocrine disruptor [55–57], in immunotoxicity [55], reproductive toxicity [58–60], neurotoxicity [61–64], cardiovascular toxicity [65], nephrotoxicity [66, 67], genotoxicity [68, 69], gastrointestinal toxicity with ulceration and hemorrhage [70–72] and it also influences organismal survival [73]. Mercury also crosses the placental barrier [74] and the blood–brain barrier and accumulates in different brain areas including the cerebellum and cerebral cortex [75, 76]. This became obvious in the tragic epidemics in Japan (Minamata disease) and in Iraq [77]. There is no single specific mechanism of toxicity ascribed to mercury, but it likely involves multiple coordinated effects on several parallel processes in the cell. Mercuric ions have a high affinity to bind to reduced sulfur atoms, especially those on endogenous thiol-containing molecules, such as GSH, cysteine, MTs, homocysteine, N-acetylcysteine (NAC), and albumin [28, 78]. It has been proposed that a mechanism of toxicity of mercury may be via binding to thiol groups, thereby damaging DNA, RNA, membrane structure, and proteins [28]. Moreover, an imbalance in the antioxidant protective mechanisms leading to oxidative stress in cells has been identified as a common factor in mercury exposure [79]. Thus, it was reported that treatment with mercury induces a dramatic rise in ROS generation leading to lipid peroxidation, protein degradation, and finally to cell death [72, 80, 81]. 3

Romero et al. Elimination of free radicals by treatment with antioxidants and or free radical scavengers leads to a reduction in the toxicity induced by mercury exposure [82, 83]. Numerous reports have documented the protective actions of melatonin in various models of oxidative stress due to its high efficacy as a free radical scavenger and indirect antioxidant [84–90]. Melatonin, as a direct free radical scavenger and indirect antioxidant, can detoxify numerous ROS including hydrogen peroxide (H2O2), hydroxyl radical (OH), peroxyl radicals (ROO), and singlet oxygen (1 O2 ), and also RNS including nitric oxide radical (NO) and peroxynitrite (ONOO) [91, 92]. Moreover; it stimulates the activities of enzymes that metabolize reactive species, thus preserving the structural and functional make up of subcellular organelles [69]. Several reports using in vivo and in vitro models showed that mercury causes genotoxicity with a delay in the cell cycle, changes in the normal pattern of chromosomal distribution during cellular division, and causes structural chromosomal aberrations which lead to the failure of DNA repair processes [68, 69, 93–95]. Mercury genotoxic is believed to be attributable to its pro-oxidative effects which damage purine–pyrimidine bases and the deoxyribose skeleton in the DNA leading to reduced cell cycle kinetics [69, 93]. These changes in cell cycle kinetics and chromosomal anomalies are prevented by supplementation with antioxidants. Melatonin has been useful in reducing the toxic effects associated with certain classes of chemotherapeutic agents, mutagens, and carcinogens, acting both as an indirect antioxidant and radical scavenger [96]. Moreover, Purohit and Rao [69] reported that melatonin or a-tocopherol alone reduces significantly the genotoxic actions induced by mercury and when given in combination, a much better amelioration is observed. Kim et al. [73] reported that 6-week-old mice treated with methyl mercury (40 mg/g) had a marked reduction in survival rate over the period 30–35 days of treatment, and when this toxic metal was co-administered with melatonin (20 mg/ mL), 100% of the mice survived, likely related to the potent antioxidative action of the indoleamine. Mercury accumulates in erythrocytes and cardiac tissues, causing detrimental effects on the cardiovascular system including a reduction in arterial blood pressure [97, 98], systemic and pulmonary vasoconstriction [99], lowering of myocardial contractility, and heart failure [100, 101]. The mechanisms behind mercury’s toxicity in the cardiovascular system are not completely understood. Cardiotoxicity induced by inorganic and organic mercury seems to be a result of impairment of hemodynamic parameters, loss of baroreceptor control, and elevated oxidative damage [65]. Detoxification of free radicals by treatment with antioxidants and or free radical scavengers has been shown to greatly reduce myocardial dysfunction induced by mercury exposure [82]. Thus, melatonin was shown to have cardioprotective potential in ameliorating mercuryinduced injury via (i) its metal-chelating activity, (ii) reducing mercury accumulation, (iii) repairing neural regulatory mechanisms of arterial blood pressure control, or (iv) increasing antioxidant enzyme activity and reducing free radical-induced cytotoxicity to myocardiocytes and endothelial cell [65]. 4

Many studies have shown that mercury exposure induces neurotoxicity due to dyshomeostasis of neurotransmitters, cytoarchitectural alterations in neural organization, increased oxidative stress, and elevated cell death [61–64]. Rao et al. [102] evaluated melatonin’s protective action induced by mercuric chloride in the cerebral and cerebellar cortices and in the brainstem. Mercury elicited the depletion of enzymatic activities such as adenosine triphosphatase (ATPase), succinate dehydrogenase (SDH), phosphorylase, alkaline phosphatase, acid phosphatase and altered glycogen, total protein, and lipid peroxidation levels in these brain areas, thereby affecting their respective functions. Co-treatment with melatonin (5 mg/kg b.w. i.p.) protected against the neural changes induced by mercury. Later, Rao and Purohit [103] documented fine structural alterations including a discontinuous myelin sheath around axons, swollen mitochondria, disfigured nuclei as well as nuclear membrane changes after the treatment of mercuric chloride; melatonin effectively reduced the enzyme and structural alterations exerted by mercury intoxication. It has also been suggested that mercury may be involved in some neurodegenerative disorders such as Alzheimer disease (AD). Hence, it was reported that mercuric chloride induces oxidative stress, amyloid-beta peptide (Ab) production, and elevated phosphorylated tau levels in neuroblastoma SH-SY5Y cells. These effects are reduced and/or totally reversed by treating the cells with melatonin. Thus, the indoleamine greatly attenuated mercury-induced oxidative stress, Ab fibrillogenesis and release, and tau hyperphosphorylation [104]. Mercury compounds also are known to affect testicular spermatogenic and steroidogenic functions in experimental animals and man [105]. Suppression of sperm motility by mercury has been reported in different mammalian species including human in association with decrements in sperm count, and altered sperm metabolism and morphology [59, 60, 93]. Mercury affects accessory sex gland function in rats and mice by producing an androgen deficiency [59]. ROS are important mediators of normal sperm function and are involved in the induction and development of sperm hyperactivation, capacitation, and the acrosome reaction [106]. But excessive production of ROS above normal levels, as mercury is capable of doing, results in lipid peroxidation and membrane damage leading to loss of sperm motility [107], inactivation of glycolytic enzymes [108], damage to the acrosomal membranes [108], and DNA oxidation, which render the sperm cell unable to fertilize the oocyte, or produce a viable pregnancy [109]. The inhibition of steroidogenesis and spermatogenesis by mercury is related to the induction of oxidative stress, lipid peroxidation, and decreasing antioxidant parameters. Melatonin when co-administered with mercury reduces oxidative stress, lipid peroxidation and recovers the androgenic production and testicular spermatogenic functions [105]. Mercury is a known endocrine disruptor, and the most often affected hormones include thyroxin, insulin, estrogen, testosterone, and adrenaline [110]. These agents are responsible for the maintenance of homeostasis, reproduction, development, and behavior [110]. When exposure occurs, mercury levels significantly increase in thyroid of animals and humans [105, 111, 112]. Mercury blocks

Melatonin for protection against metals thyroid hormone production by occupying iodine-binding sites and inhibiting the action of this hormone [105]. Mercury causes hypothyroidism, damage of thyroid RNA, autoimmune thyroiditis, and impairment of conversion of thyroxine to its active form triiodothyronine [105, 113]. Chronic intake of mercury for more than 90 days results in signs of mercury poisoning, together with decreased uptake of iodine and depression of thyroid secretion [57, 114]. Rao et al. [102] reported the mercuric chloride exposure in rats decreased thyroid activities of ATPase and succinate dehydrogenase (SDH) and reduced the generation of ATP, revealing alterations in the oxidative energy metabolism as a result of lesions in tricarboxylic acid cycle by binding with sulfhydryl groups of proteins. They also documented the existence of oxidative stress induced by inorganic mercury in the thyroid, which imposed a significant decline in levels of SOD, catalase (CAT), glutathione peroxidase (GPx), glutathione reductase (GRd), and GSH followed by elevated level of lipid peroxidation and an inhibition of thyroid function. Melatonin co-administration with mercuric chloride stimulated the antioxidative enzymes, CAT and GPx, and caused partial recovery in SDH and ATPase activities. Sener et al. [81] observed mercuric chloride toxicity at the level of the kidney, liver, and lungs; this damage was observed to be associated with increases in lipid peroxides, indicating oxidative tissue damage, as well as a rise in myeloperoxidase activity due to neutrophil infiltration, and also a significant reduction in GSH levels. Treatment with antioxidants such as melatonin protected against acute mercuric chloride toxicity by reducing of free radical production and preventing neutrophil infiltration, as well as promoting GSH synthesis. Administration of the melatonin both before and after mercuric chloride administration did not seem to afford additional protection in comparison with a single administration subsequent to mercuric chloride treatment. Mercury is an established nephrotoxicant in animals and humans where it affects the pars recta (S3 segment) of the proximal tubules [28, 66, 115]. Tubular necrosis and kidney damage, induced both in vitro and in vivo, is attributed to oxidative damage [116, 117]. As shown immunohistochemically, in the rat kidney mercuric chloride induces specific stress proteins related to mitochondrial abnormalities [118]. Stress proteins are universally conserved proteins that are a reliable index of repair in injured renal cells [67]. Melatonin reduces nephrotoxicity if administered either before or after the mercury [81, 119]. Its antioxidant effect is corroborated by reduced expression of a mitochondrial chaperone, GRP75, MT, and iNOS in rat proximal convoluted tubules [66]. In addition, maintenance of regular morphology and mitochondrial size and density in S3 segments by melatonin were shown using ultrastructure and morphometric analyses. Arsenic and melatonin The metalloid arsenic (As) is a natural environmental contaminant and a well-documented carcinogen [120] that exists mainly in two biologic oxidation states, arsenite [As (III)] and arsenate [As (V)]. The most toxicologically

potent As compounds are the trivalent forms [121] that can react with the sulfhydryl groups of proteins to inhibit many biochemical pathways. Pentavalent forms, which are less toxic, are phosphate analogues that uncouple oxidative phosphorylation. However, humans are exposed to both trivalent and pentavalent forms. Inorganic As can be either methylated to form monomethylarsonic acid [MMA (V)] or dimethylated As in dimethylarsinic acid [DMA(V)]. Complex metabolism and As biotransformation may play a pivotal role in the toxic and carcinogenic effects. For more than one decade, it has been known that methylation of inorganic As was considered the detoxification mechanism [122]. However, recent reports have shown the existence of trivalent intermediates [123], monomethylarsonous acid [MMA(III)], and dimethylarsinous acid [DMA(III)], in human urine which are more reactive and carcinogenic than pentavalent compounds. Thus, methylation of As would not be a detoxification mechanism. Currently, it has been established that As metabolism may follow two pathways, on the one hand, reduction and oxidative methylation, and on the other hand, GSH conjugation that plays a key role in both the enzymatic and nonenzymatic reduction in pentavalent arsenicals to the trivalent state (Fig. 3). Considering that As is related to diverse pathologies such as hepatic and renal disorders, cardiovascular dysfunction, neurological defects and that it has carcinogenic actions, the exact action mechanism of arsenicals is rather complicate and poorly understood. In recent years, oxidative stress has gathered strength as a likely mechanism of As toxicity [124]. During As exposure, the production of ROS alters the homeostatic balance of antioxidant defenses triggering oxidative stress [125]. Several reports have documented generation of ROS/RNS during arsenic metabolism [126, 127]. Thus, GSH depletion due the consumption induced by As metabolism may be an important step in initiating and spread of free radicals, which are likely to play an important role in the early stages of carcinogenesis. ROS generation induced by As involves pro1 duction of the superoxide anion (O O2 , OH, 2 ),  hydrogen peroxide (H2O2), ROO , and also RNS such as the NO. Furthermore, changes in antioxidant enzymes such as CAT and SOD were observed in in vitro models after As exposure [128, 129]. Reactive oxygen species are critical elements in the signal transduction pathways and transcription factor regulation. Several studies have shown that exposure to As and/ or As-induced ROS generation activate a cellular transcription factors including NF-jB, AP-1, and p53 and signaling pathways through MAP kinases [130, 131] (Fig. 4). Low doses of As increase the activity of the antioxidant enzymes including SOD, CAT, GPx, glutathione S-transferase (GST), and GRd; however, chronic exposure to this metal results in reductions in these enzyme activities. It also acts as modulator of the activity of thioredoxin reductase, heme oxygenase reductase, and NADPH oxidase [132]. Although not directly mutagenic, As is considered a comutagenic and genotoxic metal, as observed in in vitro and in laboratory animals and humans. It induces deletion mutations, oxidative DNA damage, DNA strand breaks, 5

Romero et al.

Fig. 3. Schematic diagram of arsenic metabolism.

sister chromatid exchanges, aneuploidy, transforming activity, and genomic instability [133]. Nesnow et al. [134] using a supercoiled /X174 DNAnicking assay, showed that DNA injury activity of DMA (III) is an indirect genotoxic effect mediated by ROS. The use of lM concentrations of tiron, melatonin, or trolox inhibited the DNA-nicking activities of both MMA(III) and DMA(III). However, to cause in vivo oxidative damage, the authors used high concentrations of As to promote cytotoxicity. The co-administration showed that melatonin slightly inhibits arsenite-induced in vitro urinary bladder cytotoxicity, but had no effect on the other arsenicals [135]. Perhaps in these studies, the dose of melatonin used (1000 ppm) and the route of administration chosen (oral) reduced its efficacy. When rats were injected with sodium arsenite (5.55 mg/kg b.w. 6

i.p.) for a period of 30 days, significant reductions in enzyme activities in kidney and liver were observed. Melatonin supplementation (10 mg/kg b.w. i.p.) for the 5 days prior to sacrifice reversed arsenic-induced metabolic toxicity [136]. Using the same doses of melatonin and sodium arsenite, these authors also observed changes in the antioxidant system after As exposure including reductions in the activities of SOD and CAT and suppression of the level of GSH and GRd activity in liver and kidney. Melatonin supplementation completely restored all parameters altered by As to control levels with exception of CAT [137]. Oxidative stress, after the application of sodium arsenite, also was evaluated in the nigrostriatal dopaminergic system of rat brain by Lin et al. [138]. This report demonstrated that the neuroprotective effects of melatonin

Melatonin for protection against metals

Fig 4. Arsenic-induced tumorigenesis and differentiation through cell signaling epidermal growth factor receptor (EGFR) via MAP kinases. Reactive oxygen species (ROS) generated after exposure to arsenic also activates molecules such as Ras and ERK and IjB complex. IjB complex phosphorylation leads to p65 and p50 migration to the nucleus which results in activation of pro- and anti-inflammatory cytokines which play important roles in carcinogenesis. ROS also activate nuclear factor (erythroid-derived 2)-like 2 (Nrf2), a key transcription factor that regulates the cellular antioxidant response. The oxidation of As3+ to As5+ under physiological conditions results in the formation of H2O2. Melatonin counteracts the toxic and tumorigenic effects caused by As, by (i) scavenging ROS production and stimulating antioxidative enzymes, (ii) increasing GSH levels and, (iii) modulating transcription factors such as NFjB and Nrf2.

against As-induced apoptosis involved both mitochondrial and endoplasmic reticulum processes. These data suggest that melatonin may be a good therapeutic tool to reduce oxidative stress caused by As in the central nervous system (CNS). A subsequent report evaluated the role of Asinduced oxidative stress in peripheral neurotoxicity using dorsal root ganglion explants [139]. Taking into account the pivotal role that the mitochondrial and endoplasmic reticulum pathways play in oxidative stress, melatonin likely worked via these pathways to inhibit As-induced apoptosis and oxidative stress (Fig. 4). When compared with two antioxidative thiols, GSH and NAC, melatonin was more neuroprotective than were these agents. These two studies demonstrated the neuroprotective efficacy of melatonin against As in both the central and peripheral nervous systems. Arsenic also induces apoptosis and oxidative stress in rat testes. Administration of sodium arsenite (NaAsO2) (5 mg/kg/day, intragastrically) increased the number of apoptotic germ cells and the biomarker of lipid peroxidation, MDA, while reducing SOD, CAT, and GPx activities. The concurrent treatment of rats with NaAsO2 plus

melatonin (25 mg/kg/day, i.p.) counteracted As-induced testicular apoptosis and oxidative stress [140]. The abundance of polyunsaturated fatty acids (PUFA) in the brain renders it particularly vulnerable to ROS. The role of lipid peroxidation as a major factor in Asinduced oxidative stress is well known. In an in vivo study, melatonin inhibited arsenite-mediated lipid breakdown in a concentration-dependent manner in rat brain [141]. The potential of melatonin as an antigenotoxic agent against As was evaluated by Pant and Rao [142] using the comet assay to evaluate As-induced DNA damage. They reported that melatonin protected human blood cells from the exposure of pro-oxidant actions of As. The exact molecular mechanisms of As toxicity and carcinogenesis have yet to be fully identified. Current views, in addition to espousing increased oxidative stress, also advance genetic changes and altered gene expression. Some evidence suggests that inflammation may also have a role in the As-mediated toxicity. Wang et al. [131] demonstrated that cyclooxygenase-2 (COX-2) can be a target of As-mediated toxicity in human uroepithelial cells. In 7

Romero et al.

Fig. 5. Model of cellular iron and copper homeostasis and possible interaction with melatonin (Mel). Fe3+ iron-loaded transferrin (Tf) binds to TR receptor (TfR); this complex is internalized into the endosome, where Fe3+ is reduced to Fe2+. Fe2+ is transported to the cytoplasm. The labile iron pool is defined as weakly bound iron, on average, in the +2 state. Ferritin is the iron high-capacity storage protein. Ceruloplasmin (CP) transports plasma Cu+; Cu+ enters from the blood through copper transport proteins (CTR) to the cytosol. Inside the cell, it is distributed by specific chaperones to the target protein. Copper but not iron has an effective excretion mechanism. Through the Fenton reaction these metals transform H2O2 into •OH, one of the most toxic reactive oxygen species (ROS) in nature. The imbalance of iron or copper homeostasis can increase Fe2+ or Cu+ levels, and in this way induce oxidative stress. Melatonin is a free radical scavenger, antioxidant enzyme activator and as iron chelator, inhibiting metal-mediated oxidative injury.

this study, NaAsO2 increased the COX-2 expression and activity through several modulators of the MAPK pathway (P38 and JNK) involved in cellular differentiation, inflammation, and apoptosis. Melatonin treatment (0.5 mM) reduced the generation of intracellular ROS and the expression of COX-2 mRNA induced by NaAsO2. The same research group found that low dose arsenic induced a significant rise in ATF2 (activating transcription factor and a member of AP-1) expression, which plays an essential role in the cellular stress response. Melatonin treatment and JNK or p38 inhibitors decreased significantly arsenic-induced ATF2 expression [130]. Lead and melatonin Lead is an environmental and occupational toxicant which is known to damage vital organs and suppress cellular processes [143, 144]. The main negative effects of lead are neurotoxic, genotoxic, and hematological damage [145]. Lead neurotoxicity results in behavioral changes and neurochemical alterations in neurons as a result of perturbations and disruption of main structural components of the 8

blood–brain barrier, through primary injury to astrocytes and due to secondary damage of the endothelial microvasculature [146]. Lead treatment results in a significant accumulation of this metal in all brain regions with maximal levels occurring in the hippocampus. The best-documented mechanisms of lead genotoxicity are indirect and include inhibition of DNA repair and production of free radicals [147]. Lead is known to induce hematological disturbances resulting from abnormalities in cell differentiation and hemoglobin synthesis during hematopoiesis [148]. Similar to other persistent toxic metals such as arsenic, cadmium, and mercury, lead damages cellular components at least partially via elevated levels of oxidative stress. The pathogenetic actions of lead are multifactorial as it directly interrupts the activity of enzymes, competitively inhibits absorption of important trace minerals, such as calcium and zinc, and deactivates antioxidant sulfhydryl pools [149]. Free radical-induced damage by lead is accomplished by two independent, although related, mechanisms [125]. The first involves the direct formation of ROS including 1 O2 , hydrogen peroxides, and hydroperoxides, and the second

Melatonin for protection against metals mechanism is achieved via depletion of the cellular antioxidant pool, inhibiting GRd and d-aminolevulinic acid dehydrogenase (ALAD). A direct correlation between blood lead levels, ALAD activity, and erythrocyte concentrations of MDA has been observed among workers exposed to lead. There have been several attempts to use melatonin to ameliorated lead toxicity. El-Sokkary et al. [86] investigated the neuroprotective action of melatonin against lead-induced neurotoxicity in rats. In this work, melatonin almost completely attenuated the lead-induced increase in lipid peroxidation products and restored GSH levels and SOD activity. The metal also caused severe cellular damage and reduced neuronal density in the hippocampus and striatum. Again, melatonin prevented the structural damage and maintained neuronal density. The neuroprotective effect of melatonin in lead treated animals is related to its direct radical scavenging actions and its indirect antioxidant effects [150]. In addition to its antioxidant effects, the authors discuss several other mechanisms that may be involved in the neuroprotection mediated by melatonin; these include interactions with calmodulin, blockade of rises in intracellular Ca2+, changes in gene expression and activities of antioxidant enzymes and improved efficiency of mitochondrial oxidative phosphorylation. In the cultured human neuroblastoma cell line, SH-SY5Y, exposed to low levels of lead, melatonin also restored lead-induced GSH depletion and protected against apoptosis by inhibiting caspase-3 activation [151]. Melatonin prevented oxidative DNA damage in blood lymphocytes likely due to its ability to scavenge ROS [152]. Martınez-Alfaro et al. [153] investigated the effect of melatonin on DNA damage and repair in lymphocytes of rats subchronically exposed to lead. The authors showed that low levels of lead acetate treatment induced oxidative stress, while melatonin administration reduced the toxic lead effects, but its efficacy depends on the concentration of lead to which the cells are exposed. The dose of lead acetate administered correlated with the level of lead in the blood and the extent of DNA damage in lymphocytes. Finally, administration of melatonin in conjunction with lead treatment reduced hepatic and renal toxicity in rats treated with 100 mg/kg lead for 30 days [154]. Melatonin co-treatment significantly inhibited the levels of lipid peroxidation, stimulated SOD activity and GSH concentration, restored the observed morphometric parameters, and prevented the histopathological changes in the liver and kidney. Aluminum and melatonin Aluminum is not an essential element; indeed, to date, no biologic function has been determined for aluminum. However, its presence in an organism constitutes a toxicity risk [155]. Aluminum accumulation in organs or tissues results in molecular damage or dysfunction and local concentrations of this metal usually correlate with these effects. In biologic systems, the aluminum oxidation state is Al3+. It is typically found as insoluble complexes, and therefore, its bioavailability is highly reduced.

Aluminum has long been used in industry, medicine, agriculture, and water treatment, and thus, exposure to this metal is extensive. Although aluminum is ubiquitous, it has a limited bioavailability due to its insolubility; only a small fraction of aluminum present in the diet is absorbed; most aluminum is rapidly eliminated from the body. Occupational exposure to aluminum is widespread; when the particles are inhaled, they are deposited in lungs, released into blood, and thereafter they are distributed to brain were they accumulate and exert neurotoxic effects [156]. The relationship between occupational exposure and neurobehavioral impairments is controversial; some epidemiological studies indicate there is a relationship [157, 158], while others did not find any [159]. Even though aluminum has no redox capacity in biologic systems, a number of data show that aluminum, at high concentrations, causes oxidative stress through multiple mechanisms [160]. Aluminum binds to many biologic macromolecules through interactions that may displace other biologic cations from their binding site [161]. High levels of aluminum in brain are presumably related to a number of neurodegenerative disorders such as dialysis encephalopathy, AD, and Parkinson disease (PD) early 1960s [162]. The fact that oxidative stress is associated with most neurodegenerative disorders together with the high concentrations of aluminum found in brain regions led to the increasing strong support that this metal may be related to the etiopathology of AD [155, 163]. Although aluminum is clearly involved in the etiology of AD, whether it plays a major or a minor role remains unclear [164, 165]. In recent decades, numerous studies report the pro-oxidant actions of aluminum in specific neurological areas as well as the protective role of melatonin. Esparza et al. [166] reported that after intraperitoneal administration of aluminum (5 mg/kg 8 weeks) to rats, melatonin (10 mg/kg/day 8 weeks) reduced aluminum-induced prooxidant effects in several neural areas (cerebral cortex, hippocampus, and cerebellum) compared with rats treated exclusively with aluminum. As described in detail below, findings published in a variety of reports confirmed that aluminum accumulates in several brain regions and alters oxidative stress markers such as GST, GSH, GSSG, SOD, GPx, CAT, and the levels of thiobarbituric acid reactive substances (TBARS), while melatonin displays protective effects against all aspects of aluminum-mediated damage. In aluminum-exposed rats (7 mg/kg/day), oxidative stress markers and gene expression of CuZnSOD, MnSOD, GPx, and CAT, as well as the beneficial role of melatonin (10 mg/kg/day), concurrently administered, were evaluated in hippocampus, cerebral cortex, and cerebellum. The results demonstrated that in aluminum-exposed animals, melatonin promotes the activity of the antioxidant enzymes GST, CAT, and SOD, compared with their activities in control and only aluminum-exposed animals [167, 168]. Moreover, aluminum increased significantly TBARS levels, an index of elevated lipid peroxidation; as in many other studies, melatonin greatly mitigated the degree of lipid decomposition [167, 168]. 9

Romero et al. The protective role of melatonin (10 mg/kg/day in drinking water, 6 months) in the regulation of antioxidant enzyme gene expression (CAT, GRd and SOD) in the hippocampus, cerebral cortex, and cerebellum has been evaluated in aluminum-exposed (1 mg/g of diet for 6 months) rats [169, 170]. Levels of aluminum were elevated in all animals that consumed the metal. Melatonin, when orally administered to these rats, exerted an antioxidant action by increasing mRNA levels of CAT, GRd, and SOD in hippocampus, cerebral cortex, and cerebellum [169, 170]. Unexpectedly, in these investigations, aluminum did not significantly elevated lipid peroxidation [169, 170], possibly due to the lower exposure to aluminum compared with other studies or the rather poor oral bioavailability of aluminum. Abd-Elghaffar et al. [171] carried out a study where rabbits that were orally treated with aluminum (20 mg/L in drinking water) and displayed atrophy and apoptosis of cerebral cortical and hippocampal neurons. The morphological changes, as well as the degree of lipid peroxidation and inhibition of SOD, were minimized by postadministration of melatonin (10 mg/kg b.w., 15 days), and melatonin also markedly ameliorated aluminum-induced neurotoxicity [171]. More recently, in a similar study in mice in which aluminum (3.5 mg/kg b.w., 6 weeks) and melatonin (7 mg/kg b.w., 6 weeks) were administered by intraperitoneal injection, aluminum-induced structural and oxidative damage to the medulla of rats, while melatonin, once again, mitigated this induced damage [172]. Aluminum clearly accumulates in brain regions in which damage becomes apparent in neurodegenerative disorders. Invariably, melatonin minimizes aluminum-induced damage in these regions when oxidative injury is involved. These findings make melatonin worthy of investigation as a possible therapeutic agent for neurodegenerative diseases. Biologic membranes are complex, wherein numerous molecules can be targeted by aluminum; these are phospholipids that are negatively charged, thus making these bilayer lipids a priority binding site for aluminum. Several recent studies in biologic membrane models showed that aluminum causes a dose-dependent rises in in vitro lipid peroxidation [173–176]. This has been studied in human platelet membranes where aluminum induces lipid peroxidation and where melatonin is capable to protecting against aluminum toxicity. In summary of these reports, they show that aluminum dose dependently increases the formation of lipid peroxides and melatonin dose dependently inhibits the formation of the lipid peroxides induced by aluminum [174, 175]. To test melatonin’s protective effect on aluminum-induced lipid peroxidation in rat synaptosomal membranes, Millan-Plano et al. [173] incubated AlCl3 (0–1 mM) with FeCl3 (0.1 mM) + ascorbic acid (0.1 mM). Aluminum, as previously reported [174, 177], promoted iron-initiated lipid peroxidation increased levels of both MDA and 4-hydroxyalkenal (4-HDA). Co-incubation with melatonin (0.1, 1 and 5 mM) resulted in a reduction in MDA and 4-HDA levels. Melatonin at 5 mM significantly limited MDA and 4-HDA levels to each concentration of AlCl3 tested [173]. More recently, Albendea et al. [176] confirmed these findings and, in 10

addition, showed that the highest concentration of aluminum (1 mM) tested caused significant protein oxidation, as measured by protein carbonyls. At this aluminum concentration, melatonin (1–10 mM) successfully also lowered protein carbonyl levels. Collectively, the reported results are consistent with melatonin’s protection against aluminum-induced molecular damage to a variety of molecules. The protective role of melatonin in mitochondrial and synaptosomal membranes may be due, in part, to its ability to stabilize membranes against free radicals [178–180] in addition to the benefits due to its antioxidant activities. It is well known that aluminum accumulation in the kidney is associated with alterations in the renal function [181–183]. After chronic exposure in rats, aluminum accumulates in renal tissue causing several changes in renal physiology which probably relate to the associated elevated reduction in GSH, GPx, and CAT [184, 185]. Combined administration of melatonin and aluminum significantly limited TBARS formation, normalized GSH levels and partially reversed GPx and CAT activities. However, altered renal function was only partially restored by melatonin and did not alter the accumulation of aluminum [185]. Melatonin invariably reverses lipid peroxidation [186] and partially restores the loss of CAT activity [186, 187] in aluminum-exposed rats. This metal also causes hepatic dysfunction by decreasing hepatic antioxidant enzyme activities, while pre-administration of melatonin attenuates this aluminum-mediated toxicity [188]. Using adsorptive cathodic stripping voltammetry (AdCSV), Lack et al. [189] showed that melatonin interacts with aluminum. The binding of this metal by melatonin may provide protection against aluminum-induced damage. However, in a number of other investigations, melatonin did not seem to act as an aluminum chelator [168, 169, 190]. Thus, at this stage, any role of melatonin’s protection against aluminum being a result of the ability to bind the metal remains in doubt. Chromium and melatonin Chromium is one of the most common elements in the Earth and exists in several oxidation states. Hexavalent chromium [Cr(VI)] compounds, such as chromate (CrO2 4 ) and dichromate (Cr2 O2 7 ), have long been recognized as potent respiratory toxins. Inhalation of particles containing Cr(VI) (dusts, mists, and fumes) is associated with several respiratory diseases, including lung cancer, chronic bronchitis, pulmonary fibrosis, and asthma [191–194]. Chromate predominates under physiological conditions (pH > 6), and dichromate is the predominant form under acidic conditions. Chromate closely resembles sulfate (SO2 4 ) and enters cells via an anion carrier [195], but Cr(VI) compounds also readily cross the skin [193]. Cr(VI) is not itself the toxic species. Once inside the cell, Cr(VI) is quickly reduced to Cr (III), the next stable oxidation state. Cr(III) species do not easily cross cell membranes because they are generally insoluble. Cr(VI) is reduced through reactive intermediates such as Cr(V) and (IV) to the more stable chromium (III) by cellular reductants including GSH, vitamins C and B2, and flavoenzymes [196, 197]. Moreover, this reductive

Melatonin for protection against metals process also causes the generation of ROS. The Cr (VI)/(V) redox couple serves as a cyclic electron donor in a Fentonlike reaction to produce active oxygen species, resulting in the induction of DNA damage [197, 198]. More recent studies show that Cr(VI) causes irreversible inhibition of thioredoxin reductase (TrxR) and oxidation of peroxiredoxin (Prx) and thioredoxin (Trx) [199, 200], hence increasing the oxidizing state of cells. Melatonin and its metabolites are highly effective in scavenging the highly toxic OH [201, 202]. Several studies focused on the ability of melatonin to reduce Cr-induced oxidative DNA damage in vitro. Qi et al. [203] showed that melatonin reduced DNA damage produced by CrCl3 plus H2O2. In this study, melatonin was more effective than vitamin C and E in blocking the genotoxicity of Cr (III). In another report, Qi et al. [204] compared the ability of melatonin to reduce Cr-induced oxidative DNA damage with other indoles related to melatonin; these included pinoline, NAC, 6-hydroxymelatonin, and indole3-propionic acid. The authors showed that the ability of pinoline and melatonin to reduce Cr(III)-mediated 8-OHdG formation was found to be equivalent and more effective than the other indoles tested. In these two works, the authors mention that the protective action of melatonin against Cr(III)-induced DNA damage may be related to its direct OH scavenging action and due to its ability to lower H2O2 concentrations. Moreover, the authors point to the possibility that melatonin could chelated chromium in the same manner in which it reportedly chelates other heavy metals such as aluminum, cadmium, iron, copper, and lead [40]. L opez-Burillo et al. [205] observed individual and synergistic actions of melatonin and other antioxidants (vitamin C, a-lipoic acid) on oxidative DNA damage produced by CrCl3 plus H2O2. In combination with melatonin, both vitamin C and a-lipoic acid showed synergistic actions in reducing OH produced by Cr(III)/H2O2. Susa et al. [206] examined whether melatonin has an antioxidant effect on Cr(VI)-induced DNA damage in primary cultures of rat hepatocytes. Moreover, the authors examined by electron spin resonance (ESR) spectrometry the influence of melatonin on OH formation induced by Cr(V). In this study, the authors concluded that melatonin caused a decrease in the DNA damage, cytotoxicity, and lipid peroxidation caused by Cr(VI), without affecting Cr uptake and chromium distribution in cells, possibly through its ability to increase cellular levels of vitamins C and E and to directly scavenge the OH. Iron and melatonin Iron (Fe) is an essential and bioactive element required for cellular metabolism, proliferation, and differentiation. The quantity of iron in the body is tightly associated with the control of its absorption; an overload in tissues may cause serious damage. The average human body level is 4–5 g of iron firmly bound/complexed, about a 65% to hemoglobin, 10% is a constituent of myoglobin, cytochromes, and iron-containing enzymes, and 20–30% is bound to the iron storage proteins, ferritin and hemosiderin. Ferritin is a low-affinity and high-capacity storage protein; it can store up to 4500 atoms of iron per protein macromolecule.

Transferrin is a high-affinity and low-capacity protein that it can only store two atoms of Fe(III) per molecule and this protein transports iron in the plasma. Essential processes including oxygen transport, energy production, and DNA synthesis depend on Fe-containing proteins [207]. Organisms, therefore, possess proteins that preserve iron homeostasis and maintain most of the iron sequestered with only trace amounts of the metal remaining free as nonchelated or loosely chelated iron available for catalyzing free radical reactions [208, 209]. Iron’s capacity to exchange one electron in biologic reactions is extraordinary. When one or more of its six ligand-binding sites are not tightly bound, iron engages in one electron exchange reactions with the potential of producing free radicals. It possesses incompletely filled d-orbitals and has a maximal oxidation state of 6+, but only the 2+ and 3+ states are common in biologic environments. Iron is an intrinsic ROS producer. Fe(II) is unstable and tends to react with molecular oxygen forming Fe(III) and O 2 . The electronic structure of iron and its capacity to undergo redox cycling reactions by accepting [Fe(III) + 1e ? Fe(II)] or donating [Fe(II)  1e ? Fe(III)] electron(s), places iron as a leading component in the production and metabolism of free radicals in biologic systems. However, this capacity also makes iron an essential component of cytochromes, the oxygen-binding molecules, hemoglobin and myoglobin, and many enzymes. Free radical-mediated oxidative stress in cells is one of the main causes of alterations in cellular structure and function due to iron overload [210]. A known and common source of free radicals arises from the Fenton reaction, where Fe (II) is oxidized by H2O2 to Fe(III), producing a OH. Within cells, free radical production from Fe (II) is catalytic as Fe (III) is reduced back to Fe(II) at the expense of endogenous reducing species [211–213]. Fe (II) þ H2 O2 ! Fe (III) þ  OH þ OH ðFenton reactionÞ; Fe (III) þ O 2 $ Fe (II) þ O2 ðredox reactionÞ; The net reaction (Haber–Weiss reaction);   and O 2 þ H2 O2 $ O2 þ OH þ OH

The iron-catalyzed Haber–Weiss reaction is considered to be a major mechanism by which the highly reactive  OH is generated in biologic systems [214, 215]. Nevertheless, in a lower grade, transition metal ions, such as copper, are also capable of catalyzing this reaction. Iron is clearly involved in the initiation of oxidative damage and formation of ROS, such as OH or O 2 . In addition, O combines with NO to form ONOO , a 2 RNS, which is as detrimental as OH in damaging proteins and DNA, and inhibiting DNA repair mechanisms [216].  O 2 þ NO ! ONOO

Furthermore, NO binds reversibly and high affinity to Fe (II), either free iron, iron within iron–sulfur centers, or iron within hemoproteins, under physiological or 11

Romero et al. pathophysiological conditions [217]. During cellular respiration, the interaction of NO and iron-containing enzymes has a high relevance [218]. NO binds to iNOSheme iron and inhibits its activation. The dissociation of the Fe3+-NO complex competes with its reduction to the Fe2+-NO structure to inhibit iNOS activation [219]. Fe3þ þ NO ! Fe2þ þ NO ! Fe2þ  NO or Fe3þ þ NO ! Fe3þ  NO ! Fe2þ  NO Moreover, other reactions may also take place as a consequence of OH generation, for example peroxidation of PUFA in lipids. The role of lipids in cellular membranes is not only structural but also functional; thus, any change in these molecules can lead to cell dead. If an excess of Fe (II) is present in the system, lipid hydroperoxides may undergo a Fe (II)-driven reaction [220–222]. Thus, iron is also a culprit in membrane phospholipid peroxidation; by the generation of oxygen radicals, Fe (II) induces lipid peroxidation. LH þ  OH ! L þ H2 O L þ O2 ! LOO LOO þ LH ! LOOH þ L LOOH þ Fe (II) ! Fe (III) þ  OH þ LO In summary, metal-mediated formation of free radicals may cause changes in DNA, protein and lipid structure or function and lead to changes in gene expression. As iron triggers oxidations via several reactions, protecting iron from molecular oxygen in the media is an essential event to avoid iron-mediated oxidations in biologic systems that may lead to abnormalities that affect cell or tissue functions which leads to serious cellular dysfunction or death. Obviously, a chelator molecule that binds iron would reduce reactions of free or labile iron with oxygen and its metabolites. Iron has six sites where a chelator can bind; thus, a hexadentate chelator such as desferrioxamine (DFO) completely deactivates free iron. DFO has shown high activity in Fe deprivation. However, some studies report DFO limitations; thus, its efficacy is severely limited due to its low ability to permeate biologic membranes [223–225]. Other hexadentate chelators are effective but far from being ideal therapeutic agents due to their low oral availability, toxicity, or iron donation to pathogenic organisms [226]. The ongoing therapeutic strategy to prevent metal-induced damage is the development of multifunctional iron-chelating compounds that modulate multiple disease targets [227–231]. Taking into account that antioxidants are the living organism’s major resource for protection against free radical-mediated damage, it would be reasonable to 12

develop an antioxidant therapeutic treatment line against iron damage (Fig. 5). Melatonin indirectly enhances antioxidant activity by promoting the activity of GPx [232, 233] and GRd [234] as well as mRNA levels of SOD [235]; also, GSH can act as a free radical scavenger [236]. Another reported antioxidative property of melatonin involves its diverse direct free radical scavenging activities [237–241]; the hydrophilic and lipophilic nature of melatonin may also be an important factor in this regard, allowing melatonin to move freely across all cellular barriers. These features and how iron generates free radical damage in an organism suggest that melatonin could play an important role against ironinduced damage (Fig. 5). Limson et al. [40] demonstrated, using adsorptive cathodic stripping voltammetry (ACSV), that melatonin forms in situ complexes with iron. Melatonin is able to interact with Fe (III) but not with Fe (II). Based on these data, they postulated that melatonin removes free Fe (III), thus preventing its reduction to Fe (II) avoiding free radical generation. Iron-mediated oxidative damage occurs in several neurological disorders [242–244]. Kabuto et al. [245] evaluated oxidatively destroyed tissue in rat brain homogenates after injecting FeCl3 to cause acute epileptiform discharges. After the injection, oxidative damage was evaluated by measuring levels of TBARS, which were obviously elevated. When melatonin (50–2000 lM) was co-incubated with FeCl3 (100 lM), lipid peroxidation was reduced in a concentration-dependent manner. Moreover, before rats were injected with FeCl3, pretreatment with melatonin (0.2 mmol/kg i.p.) limited TBARS concentration by 66% [245]. A quantitative study developed afterward showed that melatonin administration at pharmacological doses (5 mg/kg) for 5 days prior to and 1 day after intracortical injection of iron (ferric ammonium citrate 0.1 mM) reduced neuronal death by roughly 40% [246]. In patients with Parkinson disease, the iron content and lipid peroxidation of the brain are reportedly higher than in healthy controls [247]. Lin and Ho [248] evaluated melatonin’s neuroprotective actions on iron-induced neurodegeneration in the nigrostriatal dopaminergic system. Intranigral infusion of iron into rats caused degeneration of the nigrostriatal dopaminergic system, while coinfusion with melatonin (60 lg/lL) partially prevented iron-mediated elevation of lipid peroxidation. Additionally, in rat cortical homogenates, the in vitro activity of melatonin (0.1–4 mM) co-incubated with iron (ferrous citrate 1 lM) was also evaluated. As usual, melatonin in a dose-dependent manner overcame iron-induced lipid destruction. In a subsequent study, Chen et al. [249] used iron to induce oxidative damage in the locus coeruleus (LC) where degeneration has been documented in several neurodegenerative diseases. They observed that lipid breakdown was dose dependently elevated after iron infusion. Moreover, intraperitoneal administration of melatonin attenuated iron-induced lipid peroxidation in this nuclear group and prevented apoptosis and the reduction in locomotor activity. Studies in patients with AD revealed significantly augmented levels of iron in several brain regions [242,

Melatonin for protection against metals 250–252]. Ozcankaya and Delibas [253] reported lower melatonin levels and higher serum iron levels in patients with AD relative to these in patients with non-Alzheimer’s disease . Free radical damage was measured as MDA (the end-product of oxidative degradation of polyunsaturated fatty acids). Based on their results, both aging itself and AD promote lipid peroxidation. d-Aminolevulinic acid (ALA) is a heme synthesis precursor in some inherited and acquired porphyrias; it mainly accumulates in the liver. In the presence of Fe (II), ALA generates ROS, via metal-catalyzed oxidation to 4,5-dioxovaleric acid. OH targets DNA, resulting in the formation of stable compounds of guanidine including 8-hydroxydeoxyguanosine (8-OH-dG), a specific and sensitive molecule, useful as an oxidative DNA damage biomarker. Qi et al. [254] found melatonin at concentrations >0.1 mM to reduce ALA-induced formation of 8OH-dG in a dose-dependent manner in calf thymus DNA. Melatonin was more effective than Trolox or mannitol in this system. They postulated that the highly effective protection of melatonin may be related to its direct  OH scavenging ability as well as to its ability to detoxify the precursor H2O2 [255]. Additionally, as rather high concentrations of melatonin occur in cellular nuclei [256, 257], the findings are consistent with the reduction in nuclear DNA damage. Other studies showed that melatonin at concentrations ranging from 1 to 100 lM inhibited 8-OH-dG formation and accumulation in a dose-dependent manner [205]. Thus, melatonin provides an effective protection against metal-induced DNA damage. One investigation using rat liver homogenates compared the antioxidative effects of melatonin with other nonindole antioxidants after incubation with FeSO4 and H2O2. Lipid peroxidation was measured as MDA and 4-HDA. Melatonin (5–16 lM) reduced MDA + 4-HDA levels induced by FeSO4, but it was somewhat less efficient than vitamin E or DFO but more effective than GSH or vitamin C. Melatonin exhibited synergistic effects in association with vitamins E and C and GSH as shown by the increased efficacy in reducing lipid peroxidation in rat liver homogenates [22]. A recent study, using an experimental biliary obstruction model in rats, reported the accumulation of iron in liver and blood. Also, the presence of loosely bound iron was related to oxidative stress induced by the biliary duct obstruction. Melatonin exerts a potent effect in regulating iron metabolism; this activity is also related to its action on loosely bound iron, lipid peroxidation, and tissue injury markers (alanine aminotransferase and alkaline phosphatase) in obstructive jaundice [258]. Preeclampsia is a hypertensive disorder that is associated with pregnancy, as a consequence of an imbalance between pro-oxidant and antioxidant defenses. A recent study proved that melatonin (250 lM) acts in human placental mitochondria as an inhibitor of NADPH- and irondependent lipid peroxidation (estimated from the quantity of TBARS formed), especially when combined with vitamin C (30 lM) and/or vitamin E (25 lM). In this in vitro study, melatonin markedly reduced TBARS formation in a concentration-dependent manner. When vitamin C and vitamin E were separately or collectively co-incubated with

melatonin, a greater effect was documented. In another study of Gitto et al. [22], there was also a synergistic action between melatonin and vitamin C, which may be related to the iron-chelating ability by melatonin. The authors postulate that, melatonin in combination with ascorbate and a-tocopherol may be a worthy treatment for preeclampsia; melatonin would permit a reduction in the doses of these vitamins and also strongly enhance their antioxidant effects in the placenta [259]. Adriamycin (doxorubicin) is a widely used anticancer drug with a high clinical efficacy. Due to its diverse toxicities, its use is limited. Adriamycin extracts iron from ferritin, creating an adriamycin–iron complex, which leads to lipid peroxidation [260]. Recently, Othman et al. [261] demonstrated that melatonin ameliorated oxidative stress caused by adriamycin by regulating iron levels. In this study where melatonin (15 mg/kg) was used before and co-currently with adriamycin, the inclusion of melatonin resulted in significantly decreased levels in iron levels in plasma compared with rats treated only with adriamycin. In each of these studies, melatonin exhibited a suppressive effect against iron-induced oxidative stress. Several processes may be involved in melatonin’s protection; (i) as a free radical scavenger [237–241], (ii) as an antioxidative enzyme activator [235, 248], and (iii) as an iron chelator [40]. Based on the published data, melatonin inhibits ironinduced endogenous oxidative injury in several tissues including the brain and liver, it may be useful as a treatment for pathologies for preeclampsia, and it also regulates iron levels increased by adriamycin. The protection afforded by melatonin supports the suggestion that melatonin may be an effective therapeutic tool in iron-induced oxidative stress. Copper and melatonin Copper, with other metals, has an essential role in several physiological functions after binding to specific proteins, participating in redox reactions, aiding the transport of oxygen and storage or transport of the metal itself. Copper is the third most abundant trace element in the human body; in biologic systems, it is found as copper (I) (Cu+) and copper (II) (Cu+2). Oxidation states of copper place this metal as a useful ion for redox reactions, functioning as an electron transfer intermediate. Copper’s catalytic activity as a cofactor in redox reactions is critical for several enzymes including SOD, ascorbate oxidase, ceruloplasmin, lysyl oxidase, cytochrome c oxidase, tyrosinase, and dopamine-b-hydroxylase and is required for functions such as cellular respiration, free radical defense, neurotransmitter synthesis, and neuronal myelination [262, 263]. Its capacity to participate in one electron change reactions makes copper both essential for life and toxic because it induces oxidative damage. Ceruloplasmin (CP) is the protein responsible for the transport of 95% of plasma copper. It is synthetized in liver where it binds to copper ions. Deficiencies in CP lead to an accumulation of copper in blood and also in several tissues. Unlike iron, copper has an effective excretion mechanism. Several proteins are involved in the 13

Romero et al. intracellular transport of copper; metallochaperones are metal receptor proteins responsible for the transport of copper to its intracellular destination and they also protect this ion from reactions at a multitude of alternative binding sites [264]. Inside the cell it is distributed by specific chaperones. When in excess, metallothioneins are induced which then sequester the excess copper [265, 266] (Fig. 5). Copper can catalyze ROS formation via Fenton and Haber–Weiss chemistry [267] and can induce oxidative stress by two mechanisms: (i) copper can directly catalyze the formation of ROS via a Fenton-like means [267, 268]. The cupric ion, Cu(II), in the presence of O or 2 biologic reductants such as ascorbic acid or GSH, can be reduced to cuprous ion, Cu(I), which catalyzes the formation of the highly reactive OH through the decomposition of H2O2 via Fenton reaction [269]. (ii) Exposure to elevated levels of copper depletes GSH concentrations [270]. Disruption of copper homeostasis resulting in elevated pools of copper may contribute to a shift in redox balance toward a more oxidizing environment due to the depletion of GSH levels [271]. A reduction in GSH may enhance the cytotoxic effect of ROS and allow the metal to be more catalytically active, thus producing higher levels of ROS. On the other hand, antioxidant defenses can be directly or indirectly compromised under conditions of a copper deficiency; a loss of CP, selenium-dependent GPx, Cu/Zn-SOD, and catalase activity as well as a rise in GSH and metallothionein has been observed [272]. Cu (I) þ H2 O2 ! Cu (II) þ  OH þ OH ðFenton reactionÞ Cu (II) þ O 2 $ Cu (I) þ O2 ðredox reactionÞ Also, copper induces, via oxygen free radicals, the oxidation of bases and causes DNA damage; in both states, copper is more active than iron and induces DNA breakage [273]. Under in vitro conditions, copper exerts higher oxidative damage than iron mainly due to copper’s ability to bind proteins nonspecifically [274]. Moreover, Rae et al. [275] confirmed that free copper within the cells is limited to less than one free copper ion per cell, suggesting that cells are capable of maintaining copper in a bound state, thereby avoiding the formation of free radicals. Indeed, when iron and copper chelators were used to prevent H2O2-induced DNA damage, only iron chelators were effective, indicating that the main redox-active transition metal inside the cell is iron [276]. Copper overload due either to a high dietary intake or to transport alterations can result in many well-known disturbances, such as Wilson’s or Menkes disease. An imbalance between copper absorption and excretion can increase copper levels, and cells may not be able to maintain all the copper in the protein-bound state, mediating free radical production and oxidation of lipids, proteins, and DNA, causing impaired cellular functions and eventually cell death [263, 277]. Copper toxicity is related to its pro-oxidant activity. It is well accepted that copper can irreversibly and nonspecifically bind to thiol and amino 14

groups in proteins [278]. In addition, alterations of copper distribution are associated with neurodegenerative diseases, including AD, amyotrophic lateral sclerosis, PD, prion diseases, and Huntington’s disease [279–283]. Markers of elevated oxidative stress have been documented in a variety of tumors, possibly due to the combination of factors such as elevated active metabolism, mitochondrial mutation, cytokines, and inflammation [284]. Elevated copper levels have been shown to be directly linked to cancer progression [285]. Copper is important also for angiogenesis. Moreover, copper is known to bind to Ab via histidine and tyrosine residues [286]. Cu(II) interaction with Ab promotes its neurotoxicity which correlates with the metal reduction [Cu(I)] and with the generation of oxidative stress contributing to AD progress. The copper complex of Ab (1–42) has a highly positive reduction potential, characteristic of strongly reducing cupro-proteins. Abnormal levels of copper are also linked with diabetes [287], atherosclerosis [288], and cardiovascular disease [289]. Ab-Cu(II) þ AscH $ Ab-Cu(I) þ Asc þ Hþ Ab-Cu(II) þ Asc $ Ab-Cu(I) þ Asc Ab-Cu(I)þH2 O2 $Ab-Cu(II)þ OHþOH (Fentonreaction)

Ab-Cu(I) þ O2 $ Ab-Cu(II) þ O 2 As with other metals, some in vitro studies using electrochemical methods indicate that melatonin forms complex with copper, exhibiting a concentration-dependent affinity [40, 290]. This means that melatonin, besides acting as a potent free radical scavenger, may also bind copper preventing it from participating in free radical generation in vivo. In these studies, it was not possible to determine whether this was a simple ligation or a chelation. There is an interesting relationship in fish exposed to the copper concentrations of the water and a loss of the circadian locomotor rhythm; copper did not impede the secretion of melatonin, but the loss of locomotor activity implied that fish were not able to respond to it probably due to melatonin interacting with copper [291]. Also, Parmar and Daya [292] studied the effect of copper on pineal indoleamine synthesis and found that melatonin, when coadministered with copper, prevented N-acetyltransferase inhibition induced by copper. Parmar et al. [290] examined copper-induced lipid peroxidation in rat liver homogenates and concluded that coincubation of copper and melatonin affords protection against associated free radical damage. Wakatsuki et al. [293] determined the effects of melatonin in copper-mediated low-density lipoprotein (LDL) oxidation in normolipidemic postmenopausal women, concluding that melatonin may protect LDL against oxidation. There are other reports indicating that melatonin inhibits in vitro Cu

Melatonin for protection against metals (II)-induced LDL oxidation [294–296]. As free radicals may induce oxidative modification of LDL, a process believed to be involved in atherogenesis, the authors summarized that melatonin can inhibit oxidative modification of LDL because of its free radical scavenging effect and its interaction with copper as well. Mayo et al. [297] investigated the protective effect of melatonin, compared with other antioxidants, against oxidative protein damage induced by metal-catalyzed reactions. They concluded that melatonin was more effective and consistently protected against the structural damage caused by Cu (II)/H2O2 than other antioxidants (Trolox, ascorbate). Numerous reports have documented the protective actions of melatonin on DNA and lipids; in this article, the authors also demonstrated the effectiveness of melatonin in protecting proteins from peroxyl radicals. Zatta et al. [298] investigated the action of melatonin in free radical formation due to the interaction between b-amyloid peptides and metal ions. In this work, the authors showed that various b-amyloid peptides produce a relatively modest quantity of free radicals in the absence of metal ions. This production is greatly increased in the presence of Cu(II) and Fe(II). Furthermore, melatonin had a large inhibitory effect on free radical production. The authors concluded that melatonin may have great pharmacological potential for reducing the free radical production produced by the interaction between b-amyloid peptides and metal ions. In conclusion, the characteristic electronic configuration of copper provides its chemical properties favoring its biologic functions but also confers to copper the ability to generate excessive amounts of ROS. Copper overload toxicity due to a redox imbalance can be addressed by scavenging free radicals or chelating the excessive copper, and melatonin may fulfill these requirements, thereby providing protection against copper toxicity. Removal of excessive copper from proteins such as CuZn-SOD can have negative consequences; therefore, copper chelation therapy should be directed specifically to the site or tissue of interest. Nickel and melatonin Although a trace element for plants, some bacteria, and invertebrates (it is the central metal of ureases), nickel is toxic to humans. Nickel is not an abundant element on Earth, but both industrial and social uses, such as electroplating, and its use as anticorrosive, in batteries, jewelry, or decoration, have stimulated its use. Nickel passes into the atmosphere, water, and soil from the burning of fossil fuels, metal foundries, or tobacco smoke. Humans take in nickel by inhalation, dermal absorption, or ingestion of contaminated food. It is important to note that vegetables grown near to dense traffic roads present high concentration of nickel [299]. Nickel toxicity greatly depends on what chemical species are formed. Nickel salts are not that toxic, but soluble Ni2+ cations are highly toxic, although they are eliminated from the body easily [300]. Nickel particles target mainly the respiratory tract, leading to respiratory pathologies and cancer. Indeed, nickel species induce tumors in several animal models and sites of application [301]. The most

carcinogenic species of nickel are the disulfides and oxide dusts, nickel sulfate, nitrate, and chloride, and the vapor containing carbonyl nickel [302]. Tetracarbonyl nickel [Ni (CO)4] because of its high toxicity, penetration into cells, and persistence deserves special consideration. Regarding the biochemical behavior of nickel, it easily penetrates into cells where it binds low molecular weight proteins. Ni2+ selectively blocks T-type voltage-dependent calcium channels (VDCC) (at 50 lM), but at higher concentrations, it blocks all types of VDCC. Interestingly, at lower concentrations than those blocking VDCC, Ni2+ enters to the cell through VDCC, affecting intracellular Ca2+ homeostasis [303]. From a toxicological point of view, Ni2+ reportedly damages DNA by ROS generation derived from the catalyzed Fenton reaction, which preferentially forms H2O2, and by activating pro-inflammatory signals such as the transcription factor NF-jB. This is also related to skin hypersensitivity elicited by nickel complexes [304]. DNA damage suffers depletion of its supercoiled structure. Furthermore, DNA repair systems are disrupted by nickel, contributing to the total toxic effect through a mechanism that implicates the binding to DNA repair enzymes and the subsequent protein degradation [305]. Allergic contact dermatitis, triggered by skin exposure to Ni2+, is well accepted as being the result of the formation of organometallic complexes between nickel and electron-rich ligands [306], basically nucleophilic amino acids, generating the subsequent antigen. The oxidative stress scenario elicited by nickel is not only responsible for its carcinogenic effect, but also it is known that lipid peroxidation occurs in human plasma when nickel chloride is administered [307]. Another important target of nickel is HIF-1, as noted for other metals [308]. Nickel would displace iron from the oxygen carrier of hemoglobin, that is, hemo group, what induces a permanent hypoxia, thus activating HIF-1 [309]. Studies have considered the role of nickel in neurotoxic processes affecting cultured neurons and cell lines in relation to the protective effect of melatonin. Nickel-triggered oxidative stress and ROS generation in cortical neurons and neuro-2a neuroblastoma cells are mitigated by incubation with melatonin at millimolar concentrations [310]. Incubation of nickel chloride at concentrations from 0.125 to 1 mM produced a dose-dependent increase in ROS, measured with the fluorescent dye DCFH-DA, in both cortical neurons and neuro-2a cells; this was more apparent in the primary cultures, as ROS generation was enhanced up to threefold in the presence of 1 mM Ni2+, while neuro-2a cells only exhibited a 1.7-fold augmentation of ROS at the same concentration of Ni2+. Under similar experimental conditions, mitochondrial O con2 centrations, measured with the redox-sensitive dye MitoSOX Red, were elevated along with increasing concentrations of Ni2+ [311]. In this scenario, cell viability, measured with the method of the MTT reduction, was reduced during exposure to increasing concentrations of nickel, with a maximal reduction at 1 mM Ni2+ (by 50% in both cell cultures). Under these conditions, incubation with 1 mM melatonin dramatically reduced ROS generation, as well as mitochondrial O concentrations. 2 15

Romero et al. Melatonin prevented Ni2+-induced ROS production when [Ni2+] was between 0.125 and 0.5 mM. Ni2+-elicited ROS production was partially diminished by 1 mM melatonin in cortical neurons; ROS levels were increased by 50% over those in cells not exposed to melatonin. Better results were found in neuro-2a neuroblastoma cells where melatonin exerted a complete prevention of ROS generation in the presence of 1 mM nickel. Indeed, no significant differences were found between control (Ni2+) and sample (nickel plus melatonin) cultures. The mitochondrial O 2 concentration was reduced by approximately 50% when 0.125, 0.25, or 0.5 mM of Ni2+ was applied in the presence of 1 mM melatonin for 24 hr [311]. Pretreatment with melatonin resulted in similar results after different times of exposure to Ni2+, from 3 to 24 hr in both cell cultures [310]. As far as the loss of cell viability due to exposure to Ni2+, 1 mM melatonin, counteracted the neurotoxic actions with maximal protection in the presence of 0.5 mM of Ni2+. At higher concentrations of Ni2+ (1 mM), melatonin only rescued 50% of the cells from death for cortical neurons and neuro-2a cell cultures. Melatonin exerted the neuroprotection over a wide Ni2+ exposure times, from 3 to 24 hr. The protective actions of melatonin in these studies were presumably related to the enhancement of mitochondrial functions, for example maintaining membrane potential, increased ATP synthesis, and stabilization of mitochondrial DNA. To confirm these suppositions, confocal microscopy and flow cytometer were performed. JC-1 dye was used to assess the preservation of the mitochondrial membrane potential. Loss of red and increase in green fluorescence accounted in JC-1-charged cortical neurons treated with Ni2+ at 0.5 and 1 mM for 12 hr, and pretreatment with melatonin preserved that loss of membrane potential. In addition, ATP content was analyzed by an ATP determination kit; DNA was reduced in both cell cultures when Ni2+ was administered at concentrations from 0.125 to 1 mM; the pretreatment with melatonin at 1 mM preserved the ATP content in the presence of all nickel doses investigated [310, 311]. The assessment of the mitochondrial DNA integrity was evaluated by quantitative real-time PCR. A marked reduction in mtDNA became apparent when cortical or neuro-2a cells were exposed to nickel, at concentrations between 0.25 and 1 mM. Pretreatment with 1 mM melatonin prevented mtDNA reduction and elevating mtDNA concentration up to the level of the control situation. Nickel also provoked a marked reduction in the mtDNA transcription, which was efficiently counteracted by melatonin in the presence of all Ni2+ concentrations evaluated. In addition, 8-OHdG was used as a biomarker of mtDNA oxidative damage in neuronal cells with an enhancement of 8-OHdG-positive cells in the presence of increasing concentrations of Ni2+. Similarly, these increases were attenuated by 1 mM melatonin, being particularly efficient when Ni2+ concentrations were 0.125 and 0.25 mM [311]. Finally, the mtDNA nucleoid structure impairment elicited by Ni2+, which causes disruption of DNA functionality in terms of replication/transcription regulation, DNA content, and DNA repair systems, was abolished by pretreatment with 1 mM melatonin. However, melatonin only reduced nucleoid structures in the presence of 0.5 mM Ni2+. These 16

results confirm the higher vulnerability of mtDNA to nickel with respect to nuclear DNA. Also, mitochondria are key targets of nickel toxicity. Hence, melatonin, due to its potent antioxidant properties which included its ability to scavenge Ni2+-triggered ROS, is an important antidote against Ni2+ exposure [310, 311]. Cobalt and melatonin Cobalt is a trace element in biota, where it is presented as trivalent and divalent ions, forming a huge number of organic and inorganic salts. In nature, cobalt is found together with other metals such as copper, manganese, nickel, and arsenic. Cobalt is mobilized into the atmosphere and water doing the burning of coal and oil, exhaust from engines, and industrial waste. Industrial uses of cobalt include as a paint additive and radioligand in nuclear medicine [28]. Humans can be intoxicated with cobalt during job exposure or overmedication of cobaltcontaining drugs, such as cobalamin preparations. Clinical symptoms of cobalt intoxication include allergies [312] and blood disorders [313], among others. Cobalt fulfills an essential role in humans as an integral part of vitamin B12. Moreover, cobalt is related to the formation of thyroid hormones [314]. However, when an excess of free cobalt occurs, cytotoxic signals are triggered, mainly derived from a free radical overgeneration [28]. Also, Co(II) carcinogenicity has been demonstrated in animal studies [315]. Cobalt inhibits DNA synthesis [316] and causes DNA damage and DNA-protein cross-linking, as well as disrupting the DNA repair system [317]. It has been reported to inhibit mitosis [318] and reduce platelet aggregation [319]. High concentrations of cobalt promote various protein modulations, for example inhibition of delta-aminolevulinic acid synthase [320], activation of arginase, and induction of acylamino acid hydrolase [321]. Also, cobalt depletes neurotransmitters from several regions of the brains [322]. Mitochondrial DNA is also target for cobalt salts [323]. Global alterations in cell physiology induced by Co(II) is similar to that described for hypoxia, as the HIF-1 alpha was found to be up-regulated, as a means of preventing its ubiquitination and proteasomal degradation [324]. In terms of neurodegenerative diseases, one of the most important hypotheses to explain AD on set and progression is the disequilibrium in brain metal concentrations, mainly focused on copper and zinc [325]. Also, however, cobalt concentrations were found to be elevated in postmortem brain tissue of patients with AD [252, 326]. Co(II)-induced free radical generation seems to depend on metal chelation, as absence of GSH or histidine reduced Co(II) efficiency [327]. This finding could be the clue to explain the protective role of melatonin against cobalt-induced free radical generation, taking into account its confirmed metal-chelating properties [40], without discarding the radical scavenging feature described for melatonin. Oliveri et al. [319] noted that SH-SY5Y neuroblastoma cells exposed to increasing concentrations of cobalt chloride induced a concentration-dependent drop in GSH levels. This reduction was prevented by 1 lM

Melatonin for protection against metals melatonin. Both the antioxidant and chelating activities of melatonin were the proposed mechanism for the protective activity of the indoleamine on GSH levels. Melatonin activates the rate-limiting enzyme in GSH synthesis, that is, gamma-glutamylcysteine synthase [328]. Melatonin also preserved the viability of cells exposed to cobalt. Furthermore, 0.3 mM CoCl2 produced a 1.5-fold increase in the Aß release from SH-SY5Y cells. This rise was not observed in Co(II)-treated SH-SY5Y cells pre-incubated with melatonin [319]. The amyloidogenic feature of cobalt may be due to its inhibition of PKC, taking into account such inhibition favors amyloid precursor protein (APP) processing toward ß-secretase-addressed Aß formation. Inhibition of PKC was presumably due to competition with Ca2+, thus reducing the activity of this Ca2+-dependent enzyme. Although the exposure to metallic cobalt particles would seem to have the most direct relationship to tumor cell proliferation, soluble cobalt salts also relate to physiopathological signals involved in angiogenesis such as vascular endothelial growth factor (VEGF), which is upregulated by HIF-1, a key mediator of the hypoxia response. As mentioned above, HIF-1 exhibits enhanced activity due to exposure to cobalt [329]. Melatonin not only diminished VEGF protein and mRNA synthesis under basal conditions, but also this reduction is more pronounced in the presence of 0.1 mM CoCl2, which drastically increased both VEGF protein and mRNA production. 1 mM melatonin abolished completely the elevations in various tumor cell lines [329]. Similar observations were found in HIF-1 expression, as melatonin counteracted the CoCl2-promoted HIF-1 overexpression. Interestingly, melatonin does not affect basal levels of HIF-1. The proposed mechanism of the VEGF underexpression has its origin in the Co(II)-induced HIF-1 inhibition. Vanadium and melatonin Vanadium is trace element essential for many living organisms, including humans. It plays several roles in biota, including as counterions of proteins, DNA, and RNA. Sufficient doses of vanadium for humans are very small, so environmental or labor exposure to vanadium species easily leads to an exceeded concentration causing toxicity [330]. Vanadium is mobilized and gets to natural beds and the atmosphere from the combustion products of oils, dusts, and fumes. Accumulation of vanadium is mainly found in liver, kidney, and bones [28]. Vanadium compounds are used in chemotherapy [331], due to their inhibitory action of xenobiotic enzymes and cancer cell metastatic potential. Other uses of vanadium compounds are as a diabetes therapy, as a contraceptive and as nutritional complement [332]. The principal vanadium species is pentavalent vanadate, which can be reduced by hydrogenases to vanadyl (IV) species; when this happens, NADPH or GSH levels are lost and H2O2 is formed. Vanadate species are capable of generating the OH and other ROS. Vanadium-induced ROS generation triggers several signaling pathways, influenced by proteins such as AP-1, MEK-1, and ERK-1. [333]. Further oxidative stress scenarios can be achieved

from its demonstrated inhibition of tyrosine phosphatases [28], and this causes loss of tyrosine phosphatase substrates, for example Na+/K+ ATPase and Ca2+-dependent ATPases. Final consequences of these activations are huge, but DNA damage is the most obvious. Like cobalt, vanadate species up-regulate the HIF-1 protein, inducing its expression through the PI3K/Akt pathway [334]. Vanadyl (IV) compounds also present free radical-derived toxicity, and this oxidative stress signal leads to programmed cell death [335]. In an isolated pulmonary arterial ring preparation, designed to measure the contractile effect of vanadate, presumably due to H2O2 overexpression, this anion evoked an inotropic response at 0.1–1 mM concentration that was efficiently annulled by 1 mM melatonin. Vanadium-elicited NADH oxidation was reduced by melatonin in a concentration-dependent fashion, as increasing concentrations of melatonin reduced the oxidation of NADH up to a 75% [336]. As previously noted, therapeutic use of organic chelated vanadium compounds includes as a diabetes treatment. This is because vanadium has been found to possess insulin-mimetic properties, as well as enhancing the effects of insulin when administered orally [337]. However, overuse of vanadium-derived drugs to treat diabetes leads to toxicological phenomena, preferentially at the kidney level. Administration of organic chelated vanadium resulted in a decrease in the mitochondrial membrane potential, while pre-incubation with melatonin attenuated this decrease. Furthermore, vanadium-generated hydroxylated free radical species in renal tubules were reduced in preparations including melatonin. In general, co-administration of melatonin with the vanadium-derived drugs produced a reduction in serum glucose levels indicating that melatonin can have a beneficial effect in patients with diabetes treated with vanadium compounds, mainly because of its protective role on renal function [337]. Molybdenum and melatonin Molybdenum presents different oxidation states with an easy redox transfer. It occurs on Earth preferentially as a sulfured mineral (molybdenite, MoS2). Industrial uses of molybdenum include its uses in the manufacture of hightemperature-resistant alloys, catalysts, lubricants, and dyes [299]. Molybdenum is a trace element for humans and it is necessary in very low doses. It is a key component of several enzymes referred to as molybdenum hydroxylases or molybdoenzymes, which participate in the biotransformation of xenobiotics [338], that is, aldehyde oxidase, sulfite oxidase, xanthine dehydrogenase, and xanthine oxidase. It also plays a role in the metabolism of purine. The molybdoenzymes contain FAD, two iron–sulfur centers, and hexavalent molybdenum in the form of a pterin cofactor. Thus, molybdenum oxidizes substrates and is reoxidized by molecular oxygen through a mechanism involving iron centers [299]. Overexposure to molybdenum can lead to toxicity scenarios. High concentrations of molybdenum induce the molybdoenzyme xanthine oxidase overexpression. Taking 17

Romero et al. into account they use molecular oxygen to reoxidize the molybdenum-centered cofactor, their overexpression leads to a higher generation of ROS, for example O and 2 H2O2. Hence, molybdenum toxicity is well correlated with the overactivity of xanthine oxidase and related enzymes, that is, during ischemia–reperfusion injuries in heart, skeletal muscle, and brain [339]. Sulfite oxidase, which reportedly does not produce ROS, has been linked to the some neurological defects [339]. Chronic exposure to molybdenum provokes high uric acid levels, loss of appetite, diarrhea, anemia, and a slow growth rate. High molybdenum density areas have been related to the appearance of gouty diseases [340]. In summary, molybdenum toxicity is indirectly related to ROS generation, as molybdoenzymes overexpression seems to be responsible of the oxidative stress-derived injury when excessive molybdenum is present. Only a few contributions discuss the role of melatonin in the modulation such as molybdenum-triggered damage [337]. The reason for this lack of work in this area may be the tricky chelation of the high valence molybdenum species present in biota, for example molybdate (MoO2 4 ), by potential ligands such as melatonin. It is accepted that ingestion of molybdenum, and before incorporation as a center metal of molybdenum hydroxylases, molybdenum is retained in the liver linked to a nonprotein cofactor bound to the mitochondrial outer membrane. In this state, melatonin, by chelating molybdenum, could compete with the transferring apoenzyme. However, the potential organometallic complex with melatonin does not seem to be possible because of spatial and electronic issues. Thus, melatonin may not be capable of exerting protective actions against molybdenum-induced toxicity. Exploring the chelating properties of melatonin: A proposed hypothesis Melatonin is a nonenzymatic antioxidant (namely antioxidant H) that is able to counteract the oxidative actions of metals by two major means, that is, scavenging the generated ROS and capturing such metals to form chelates, thus deactivating their ability to trigger oxidative stress processes [341]. Physicochemical and theoretical studies have focused on the antioxidant properties of melatonin, extensively reviewed by Galano et al. [342]. Radical scavenging drives the formation of several metabolites including cyclic 3-hydroxymelatonin, N1-acetyl-N2-formyl-5-methoxykynuramine (AFMK), and N-acetyl-5-methoxykynuramine. By contrast, the corroborated chelating properties of melatonin [40] are not that simple to explain from a chemical point of view, as indole nitrogen does not have the ability to complex electrophile elements, taking into account its lone pair contributes to the aromaticity of the heterocyclic nucleus of melatonin. It is worthwhile mentioning that any metal, proposed to be chelated by melatonin, would present its own requirements to achieve such a chelating complex, following organometallic rules for the ligand–metal coordination, namely the 18 electrons rule (16 electrons for those called ‘platinum metals’, e.g., Ni2+ complexes) and the generation of tetrahedral, octahedral, or square planar-like 18

structures. For all of these reasons, a high possibility exists that the p-electrons of the benzene-fused ring are recruited to smoothly bind the metal, in a g6-hapticity fashion, depending on the number of electrons necessary to form a stable organometallic complex. Thus, a hypothetical representation could be as schematized in Fig. 2.

Concluding remarks This review summarizes the impact of metal-induced free radical formation on cells and how melatonin may counteract its toxic effects. Metals induce ROS/RNS generation which triggers epigenetic changes, abnormal cell signaling, uncontrolled cell growth, initiation of cellular injury, and the stimulation of inflammatory processes. Melatonin constitutes a valuable protector versus metal-induced damage due to its multiple properties: (i) direct free radical scavenger, (ii) increasing activity and expression of antioxidant enzymes, (iii) high lipophilicity which makes its easy transport across cellular membranes, (iv) chelating activity, and (v) its low toxicity. Certainly, after the analysis of the studies carried out in animal models and in vitro systems performed, it can be concluded that melatonin is a promising therapeutic tool for metal-related damage. However, further studies are necessary to confirm the usefulness of melatonin against metal-induced toxicity. Nevertheless, the realization of clinical trials on the possible usefulness of melatonin reducing metal-induced toxicity may result in promising data for several diseases that involve metals in their physiopathology.

Acknowledgements C. de los Rios thanks IS Carlos III for research contract under Miguel Servet Program. We thank Eva M. GarcıaFrutos (Instituto de Ciencia de Materiales, CSIC, Madrid, Spain) for co-crystallization of melatonin with selected metal.

References 1. LEONARD SS, HARRIS GK, SHI X. Metal-induced oxidative stress and signal transduction. Free Radic Biol Med 2004; 37:1921–1942. 2. BAL W, KASPRZAK KS. Induction of oxidative DNA damage by carcinogenic metals. Toxicol Lett 2002; 127:55–62. 3. CHEN F, SHI X. Intracellular signal transduction of cells in response to carcinogenic metals. Crit Rev Oncol Hematol 2002; 42:105–121. 4. REITER RJ, TAN D-X, FUENTES-BROTO L. Melatonin: a multitasking molecule. Prog Brain Res 2010; 181:127–151.  ~ -OTALORA 5. ESCAMES G, OZTURK G, BANO B et al. Exercise and melatonin in humans: reciprocal benefits. J Pineal Res 2012; 52:1–11. 6. HARDELAND R, MADRID JA, TAN DX et al. Melatonin, the circadian multioscillator system and health: the need for detailed analyses of peripheral melatonin signaling. J Pineal Res 2012; 52:139–166. 7. CARDINALI DP, SRINIVASAN V, BRZEZINSKI A et al. Melatonin and its analogs in insomnia and depression. J Pineal Res 2012; 52:365–375.

Melatonin for protection against metals 8. CARBAJO-PESCADOR S, STEINMETZ C, KASHYAP A et al. Melatonin induces transcriptional regulation of Bim by FoxO3a in HepG2 cells. Br J Cancer 2013; 108:442–449. 9. MAURIZ JL, COLLADO PS, VENEROSO C et al. A review of the molecular aspects of melatonin’s anti-inflammatory actions: recent insights and new perspectives. J Pineal Res 2013; 54:1–14. 10. LIU J, HUANG F, HE H-W. Melatonin effects on hard tissues: bone and tooth. Int J Mol Sci 2013; 14:10063–10074.  11. CALVO JR, GONZALEZ -YANES C, MALDONADO M. The role of melatonin in the cells of the innate immunity: a review. J Pineal Res 2013; 55:103–120. 12. REITER RJ, TAN D-X, LEON J et al. When melatonin gets on your nerves: its beneficial actions in experimental models of stroke. Exp Biol Med 2005; 230:104–117. 13. KILIC U, YILMAZ B, UGUR M et al. Evidence that membrane-bound G protein-coupled melatonin receptors MT1 and MT2 are not involved in the neuroprotective effects of melatonin in focal cerebral ischemia. J Pineal Res 2012; 52:228–235. 14. ABE M, REITER RJ, ORHII PB et al. Inhibitory effect of melatonin on cataract formation in newborn rats: evidence for an antioxidative role for melatonin. J Pineal Res 1994; 17:94–100. 15. GAO C, HAN HB, TIAN XZ et al. Melatonin promotes embryonic development and reduces reactive oxygen species in vitrified mouse 2-cell embryos. J Pineal Res 2012; 52:305–311. 16. DUAN W, YANG Y, YI W et al. New role of JAK2/STAT3 signaling in endothelial cell oxidative stress injury and protective effect of melatonin. PLoS ONE 2013; 8:e57941. 17. GALANO A, TAN DX, REITER RJ. On the free radical scavenging activities of melatonin’s metabolites, AFMK and AMK. J Pineal Res 2013; 54:245–257. 18. PABLOS MI, AGAPITO MT, GUTIERREZ R et al. Melatonin stimulates the activity of the detoxifying enzyme glutathione peroxidase in several tissues of chicks. J Pineal Res 1995; 19:111–115. 19. RODRIGUEZ C, MAYO JC, SAINZ RM et al. Regulation of antioxidant enzymes: a significant role for melatonin. J Pineal Res 2004; 36:1–9.  20. FISCHER TW, KLESZCZYNSKI K, HARDKOP LH et al. Melatonin enhances antioxidative enzyme gene expression (CAT, GPx, SOD), prevents their UVR-induced depletion, and protects against the formation of DNA damage (8-hydroxy-20 -deoxyguanosine) in ex vivo human skin. J Pineal Res 2013; 54:303–312. ~ -CASTROVIEJO D, COTO-MONTES 21. ROSALES-CORRAL SA, ACUNA A et al. Alzheimer’s disease: pathological mechanisms and the beneficial role of melatonin. J Pineal Res 2012; 52: 167–202. 22. GITTO E, TAN DX, REITER RJ et al. Individual and synergistic antioxidative actions of melatonin: studies with vitamin E, vitamin C, glutathione and desferrioxamine (desferoxamine) in rat liver homogenates. J Pharm Pharmacol 2001; 53:1393–1401. 23. LYSENG-WILLIAMSON KA. Melatonin prolonged release. Drugs Aging 2012; 29:911–923. 24. KOTLARCZYK MP, LASSILA HC, O’NEIL CK et al. Melatonin osteoporosis prevention study (MOPS): a randomized, double-blind, placebo-controlled study examining the effects of melatonin on bone health and quality of life in perimenopausal women. J Pineal Res 2012; 52:414–426.

25. KIM KJ, CHOI JS, KANG I et al. Melatonin suppresses tumor progression by reducing angiogenesis stimulated by HIF-1 in a mouse tumor model. J Pineal Res 2013; 54: 264–270. 26. DOMINGUEZ-RODRIGUEZ A, ABREU-GONZALEZ P, ARROYOUCAR E et al. Global left ventricular longitudinal strain is associated with decreased melatonin levels in patients with acute myocardial infarction: a two-dimensional speckle tracking study. Biomarkers 2013; 18:310–313. 27. HARDELAND R. Melatonin and the theories of aging: a critical appraisal of melatonin’s role in antiaging mechanisms. J Pineal Res 2013; 55:325–356. 28. VALKO M, MORRIS H, CRONIN MT. Metals, toxicity and oxidative stress. Curr Med Chem 2005; 12:1161–1208. 29. THEVENOD F, LEE WK. Cadmium and cellular signaling cascades: interactions between cell death and survival pathways. Arch Toxicol 2013; 87:1743–1786. 30. HARTWIG A. Cadmium and cancer. Met Ions Life Sci 2013; 11:491–507. 31. THEVENOD F. Cadmium and cellular signaling cascades: to be or not to be? Toxicol Appl Pharmacol 2009; 238: 221–239. 32. HUANG YH, SHIH CM, HUANG CJ et al. Effects of cadmium on structure and enzymatic activity of Cu, Zn-SOD and oxidative status in neural cells. J Cell Biochem 2006; 98:577–589. 33. CASALINO E, SBLANO C, CALZARETTI G et al. Acute cadmium intoxication induces alpha-class glutathione S-transferase protein synthesis and enzyme activity in rat liver. Toxicology 2006; 217:240–245. 34. IKEDIOBI CO, BADISA VL, AYUK-TAKEM LT et al. Response of antioxidant enzymes and redox metabolites to cadmiuminduced oxidative stress in CRL-1439 normal rat liver cells. Int J Mol Med 2004; 14:87–92. 35. BEDARD K, KRAUSE KH. The NOX family of ROS-generating NADPH oxidases: physiology and pathophysiology. Physiol Rev 2007; 87:245–313. 36. KIM CY, LEE MJ, LEE SM et al. Effect of melatonin on cadmium-induced hepatotoxicity in male Sprague-Dawley rats. Tohoku J Exp Med 1998; 186:205–213. 37. KARBOWNIK M, GITTO E, LEWINSKI A et al. Induction of lipid peroxidation in hamster organs by the carcinogen cadmium: melioration by melatonin. Cell Biol Toxicol 2001; 17:33–40. 38. EYBL V, KOTYZOVA D, KOUTENSKY J. Comparative study of natural antioxidants – curcumin, resveratrol and melatonin – in cadmium-induced oxidative damage in mice. Toxicology 2006; 225:150–156. 39. CHWELATIUK E, WLOSTOWSKI T, KRASOWSKA A et al. The effect of orally administered melatonin on tissue accumulation and toxicity of cadmium in mice. J Trace Elem Med Biol 2006; 19:259–265. 40. LIMSON J, NYOKONG T, DAYA S. The interaction of melatonin and its precursors with aluminium, cadmium, copper, iron, lead, and zinc: an adsorptive voltammetric study. J Pineal Res 1998; 24:15–21. 41. KONAR V, KARA H, YILMAZ M et al. Effects of selenium and vitamin E, in addition to melatonin, against oxidative stress caused by cadmium in rats. Biol Trace Elem Res 2007; 118:131–137. 42. KARA H, CEVIK A, KONAR V et al. Protective effects of antioxidants against cadmium-induced oxidative damage in rat testes. Biol Trace Elem Res 2007; 120:205–211.

19

Romero et al. 43. KARA H, CEVIK A, KONAR V et al. Effects of selenium with vitamin E and melatonin on cadmium-induced oxidative damage in rat liver and kidneys. Biol Trace Elem Res 2008; 125:236–244. 44. POLIANDRI AH, ESQUIFINO AI, CANO P et al. In vivo protective effect of melatonin on cadmium-induced changes in redox balance and gene expression in rat hypothalamus and anterior pituitary. J Pineal Res 2006; 41:238–246. 45. MILER EA, NUDLER SI, QUINTEROS FA et al. Cadmium induced-oxidative stress in pituitary gland is reversed by removing the contamination source. Hum Exp Toxicol 2010; 29:873–880. 46. JIMENEZ-ORTEGA V, CANO-BARQUILLA P, SCACCHI PA et al. Cadmium-induced disruption in 24-h expression of clock and redox enzyme genes in rat medial basal hypothalamus: prevention by melatonin. Front Neurol 2011; 2:13. 47. JIMENEZ-ORTEGA V, CANO BARQUILLA P, FERNANDEZMATEOS P et al. Cadmium as an endocrine disruptor: correlation with anterior pituitary redox and circadian clock mechanisms and prevention by melatonin. Free Radic Biol Med 2012; 53:2287–2297. 48. KLAASSEN CD, LIU J, DIWAN BA. Metallothionein protection of cadmium toxicity. Toxicol Appl Pharmacol 2009; 238:215–220. 49. ALONSO-GONZALEZ C, MEDIAVILLA D, MARTINEZ-CAMPA C et al. Melatonin modulates the cadmium-induced expression of MT-2 and MT-1 metallothioneins in three lines of human tumor cells (MCF-7, MDA-MB-231 and HeLa). Toxicol Lett 2008; 181:190–195. 50. EL-SOKKARY GH, NAFADY AA, SHABASH EH. Melatonin ameliorates cadmium-induced oxidative damage and morphological changes in the kidney of rat. Open Neuroendocrinol J 2009; 2:1–9. 51. EL-SOKKARY GH, NAFADY AA, SHABASH EH. Melatonin administration ameliorates cadmium-induced oxidative stress and morphological changes in the liver of rat. Ecotoxicol Environ Saf 2010; 73:456–463. 52. JI YL, WANG H, MENG C et al. Melatonin alleviates cadmium-induced cellular stress and germ cell apoptosis in testes. J Pineal Res 2012; 52:71–79. 53. MUKHERJEE R, BANERJEE S, JOSHI N et al. A combination of melatonin and alpha lipoic acid has greater cardioprotective effect than either of them singly against cadmiuminduced oxidative damage. Cardiovasc Toxicol 2011; 11: 78–88. 54. FITZGERALD WF, CLARKSON TW. Mercury and monomethylmercury: present and future concerns. Environ Health Perspect 1991; 96:159–166. 55. ANDO T, WAKISAKA I, YANAGIHASHI T et al. [Mercury concentration in gray hair]. Nihon Eiseigaku Zasshi 1989; 43:1063–1068. 56. KAWADA J, NISHIDA M, YOSHIMURA Y et al. Effects of organic and inorganic mercurials on thyroidal functions. J Pharmacobiodyn 1980; 3:149–159. 57. NISHIDA M, YAMAMOTO T, YOSHIMURA Y et al. Subacute toxicity of methylmercuric chloride and mercuric chloride on mouse thyroid. J Pharmacobiodyn 1986; 9:331–338. 58. KHAN AT, ATKINSON A, GRAHAM TC et al. Effects of inorganic mercury on reproductive performance of mice. Food Chem Toxicol 2004; 42:571–577. 59. RAO MV. Histophysiological changes of sex organs in methylmercury intoxicated mice. Endocrinol Exp 1989; 23:55–62.

20

60. MOHAMED MK, BURBACHER TM, MOTTET NK. Effects of methyl mercury on testicular functions in Macaca fascicularis monkeys. Pharmacol Toxicol 1987; 60:29–36. 61. CECCATELLI S, DARE E, MOORS M. Methylmercury-induced neurotoxicity and apoptosis. Chem Biol Interact 2010; 188:301–308. 62. BASU N, SCHEUHAMMER AM, EVANS RD et al. Cholinesterase and monoamine oxidase activity in relation to mercury levels in the cerebral cortex of wild river otters. Hum Exp Toxicol 2007; 26:213–220. 63. ASCHNER M, SYVERSEN T, SOUZA DO et al. Involvement of glutamate and reactive oxygen species in methylmercury neurotoxicity. Braz J Med Biol Res 2007; 40:285–291. 64. MORENO-FUENMAYOR H, BORJAS L, ARRIETA A et al. Plasma excitatory amino acids in autism. Invest Clin 1996; 37:113–128. 65. JINDAL M, GARG GR, MEDIRATTA PK et al. Protective role of melatonin in myocardial oxidative damage induced by mercury in murine model. Hum Exp Toxicol 2011; 30:1489–1500. 66. STACCHIOTTI A, RICCI F, REZZANI R et al. Tubular stress proteins and nitric oxide synthase expression in rat kidney exposed to mercuric chloride and melatonin. J Histochem Cytochem 2006; 54:1149–1157. 67. Van De WATER B, De GRAAUW M, Le DEVEDEC S et al. Cellular stress responses and molecular mechanisms of nephrotoxicity. Toxicol Lett 2006; 162:83–93. 68. AKIYAMA M, OSHIMA H, NAKAMURA M. Genotoxicity of mercury used in chromosome aberration tests. Toxicol In Vitro 2001; 15:463–467. 69. PUROHIT AR, RAO MV. Mitigative role of melatonin and alpha-tocopherol against mercury-induced genotoxicity. Drug Chem Toxicol 2014; 37:221–226. 70. STACEY NH, KAPPUS H. Cellular toxicity and lipid peroxidation in response to mercury. Toxicol Appl Pharmacol 1982; 63:29–35. 71. LUND BO, MILLER DM, WOODS JS. Studies on Hg(II)induced H2O2 formation and oxidative stress in vivo and in vitro in rat kidney mitochondria. Biochem Pharmacol 1993; 45:2017–2024. 72. MAHBOOB M, SHIREEN KF, ATKINSON A et al. Lipid peroxidation and antioxidant enzyme activity in different organs of mice exposed to low level of mercury. J Environ Sci Health B 2001; 36:687–697. 73. KIM CY, NAKAI K, KAMEO S et al. Protective effect of melatonin on methylmercury-Induced mortality in mice. Tohoku J Exp Med 2000; 191:241–246. 74. CLARKSON TW. Molecular and ionic mimicry of toxic metals. Annu Rev Pharmacol Toxicol 1993; 33:545–571. 75. ERIE JC, BUTZ JA, GOOD JA et al. Heavy metal concentrations in human eyes. Am J Ophthalmol 2005; 139:888–893. 76. MOTTET NK, VAHTER ME, CHARLESTON JS et al. Metabolism of methylmercury in the brain and its toxicological significance. Met Ions Biol Syst 1997; 34:371–403. 77. WATANABE C, SATOH H. Evolution of our understanding of methylmercury as a health threat. Environ Health Perspect 1996; 104(Suppl 2):367–379. 78. HULTBERG B, ANDERSSON A, ISAKSSON A. Interaction of metals and thiols in cell damage and glutathione distribution: potentiation of mercury toxicity by dithiothreitol. Toxicology 2001; 156:93–100. 79. HUSSAIN S, RODGERS DA, DUHART HM et al. Mercuric chloride-induced reactive oxygen species and its effect on

Melatonin for protection against metals

80.

81.

82.

83.

84. 85.

86.

87.

88.

89.

90.

91.

92.

93.

94.

95.

96.

97.

antioxidant enzymes in different regions of rat brain. J Environ Sci Health B 1997; 32:395–409. DRAPER GJ, SANDERS BM, LENNOX EL et al. Patterns of childhood cancer among siblings. Br J Cancer 1996; 74:152–158. SENER G, SEHIRLI AO, AYANOGLU-DULGER G. Melatonin protects against mercury(II)-induced oxidative tissue damage in rats. Pharmacol Toxicol 2003; 93:290–296. SENER G, SEHIRLI AO, AYANOGLU-DULGER G. Protective effects of melatonin, vitamin E and N-acetylcysteine against acetaminophen toxicity in mice: a comparative study. J Pineal Res 2003; 35:61–68. KUKREJA RC, HESS ML. The oxygen free radical system: from equations through membrane-protein interactions to cardiovascular injury and protection. Cardiovasc Res 1992; 26:641–655. REITER RJ. Oxidative damage in the central nervous system: protection by melatonin. Prog Neurobiol 1998; 56:359–384. LEE YM, CHEN HR, HSIAO G et al. Protective effects of melatonin on myocardial ischemia/reperfusion injury in vivo. J Pineal Res 2002; 33:72–80. EL-SOKKARY GH, KAMEL ES, REITER RJ. Prophylactic effect of melatonin in reducing lead-induced neurotoxicity in the rat. Cell Mol Biol Lett 2003; 8:461–470. RAO MV, GANGADHARAN B. Antioxidative potential of melatonin against mercury induced intoxication in spermatozoa in vitro. Toxicol In Vitro 2008; 22:935–942. REITER RJ, PAREDES SD, MANCHESTER LC et al. Reducing oxidative/nitrosative stress: a newly-discovered genre for melatonin. Crit Rev Biochem Mol Biol 2009; 44:175–200. PARK S, LEE SK, PARK K et al. Beneficial effects of endogenous and exogenous melatonin on neural reconstruction and functional recovery in an animal model of spinal cord injury. J Pineal Res 2012; 52:107–119.  KLESZCZYNSKI K, TUKAJ S, KRUSE N et al. Melatonin prevents ultraviolet radiation-induced alterations in plasma membrane potential and intracellular pH in human keratinocytes. J Pineal Res 2013; 54:89–99. TAN DX, HARDELAND R, MANCHESTER LC et al. Mechanistic and comparative studies of melatonin and classic antioxidants in terms of their interactions with the ABTS cation radical. J Pineal Res 2003; 34:249–259. REITER RJ, TAN DX, MANCHESTER LC et al. Biochemical reactivity of melatonin with reactive oxygen and nitrogen species: a review of the evidence. Cell Biochem Biophys 2001; 34:237–256. RAO MV, SHARMA PS. Protective effect of vitamin E against mercuric chloride reproductive toxicity in male mice. Reprod Toxicol 2001; 15:705–712. CEBULSKA-WASILEWSKA A, PANEK A, ZABINSKI Z et al. Occupational exposure to mercury vapour on genotoxicity and DNA repair. Mutat Res 2005; 586:102–114. RAO MV, CHINOY NJ, SUTHAR MB et al. Role of ascorbic acid on mercuric chloride-induced genotoxicity in human blood cultures. Toxicol In Vitro 2001; 15:649–654. LIALIARIS T, LYRATZOPOULOS E, PAPACHRISTOU F et al. Supplementation of melatonin protects human lymphocytes in vitro from the genotoxic activity of melphalan. Mutagenesis 2008; 23:347–354. RHEE HM, CHOI BH. Hemodynamic and electrophysiological effects of mercury in intact anesthetized rabbits and in isolated perfused hearts. Exp Mol Pathol 1989; 50:281–290.

98. ROSSONI LV, AMARAL SM, VASSALLO PF et al. Effects of mercury on the arterial blood pressure of anesthetized rats. Braz J Med Biol Res 1999; 32:989–997. 99. Da CUNHA V, SOUZA HP, ROSSONI LV et al. Effects of mercury on the isolated perfused rat tail vascular bed are endothelium-dependent. Arch Environ Contam Toxicol 2000; 39:124–130. 100. OLIVEIRA EM, VASSALLO DV, SARKIS JJ et al. Mercury effects on the contractile activity of isolated heart muscle. Toxicol Appl Pharmacol 1994; 128:86–91. 101. EL-SHENAWY SM, HASSAN NS. Comparative evaluation of the protective effect of selenium and garlic against liver and kidney damage induced by mercury chloride in the rats. Pharmacol Rep 2008; 60:199–208. 102. RAO MV, PUROHIT A, PATEL T. Melatonin protection on mercury-exerted brain toxicity in the rat. Drug Chem Toxicol 2010; 33:209–216. 103. RAO MV, PUROHIT AR. Neuroprotection by melatonin on mercury induced toxicity in the rat brain. Pharmacol Pharm 2011; 2:375–385. 104. OLIVIERI G, BRACK C, MULLER-SPAHN F et al. Mercury induces cell cytotoxicity and oxidative stress and increases beta-amyloid secretion and tau phosphorylation in SHSY5Y neuroblastoma cells. J Neurochem 2000; 74: 231–236. 105. RAO MV, CHHUNCHHA B. Protective role of melatonin against the mercury induced oxidative stress in the rat thyroid. Food Chem Toxicol 2010; 48:7–10. 106. AITKEN RJ. Free radicals, lipid peroxidation and sperm function. Reprod Fertil Dev 1995; 7:659–668. 107. ALVAREZ JG, STOREY BT. Spontaneous lipid peroxidation in rabbit epididymal spermatozoa: its effect on sperm motility. Biol Reprod 1982; 27:1102–1108. 108. ALVAREZ JG, STOREY BT. Assessment of cell damage caused by spontaneous lipid peroxidation in rabbit spermatozoa. Biol Reprod 1984; 30:323–331. 109. GIL-GUZMAN E, OLLERO M, LOPEZ MC et al. Differential production of reactive oxygen species by subsets of human spermatozoa at different stages of maturation. Hum Reprod 2001; 16:1922–1930. 110. ZHU X, KUSAKA Y, SATO K et al. The endocrine disruptive effects of mercury. Environ Health Prev Med 2000; 4: 174–183. 111. KOSTA L, BYRNE AR, ZELENKO V. Correlation between selenium and mercury in man following exposure to inorganic mercury. Nature 1975; 254:238–239. 112. NYLANDER M, FRIBERG L, EGGLESTON D et al. Mercury accumulation in tissues from dental staff and controls in relation to exposure. Swed Dent J 1989; 13:235–243. 113. GHOSH N, BHATTACHARYA S. Thyrotoxicity of the chlorides of cadmium and mercury in rabbit. Biomed Environ Sci 1992; 5:236–240. 114. BARREGARD L, LINDSTEDT G, SCHUTZ A et al. Endocrine function in mercury exposed chloralkali workers. Occup Environ Med 1994; 51:536–540. 115. CLARKSON TW. The toxicology of mercury. Crit Rev Clin Lab Sci 1997; 34:369–403. 116. ALEO MF, MORANDINI F, BETTONI F et al. Antioxidant potential and gap junction-mediated intercellular communication as early biological markers of mercuric chloride toxicity in the MDCK cell line. Toxicol In Vitro 2002; 16: 457–465.

21

Romero et al. 117. SHIMOJO N, KUMAGAI Y, NAGAFUNE J. Difference between kidney and liver in decreased manganese superoxide dismutase activity caused by exposure of mice to mercuric chloride. Arch Toxicol 2002; 76:383–387. 118. STACCHIOTTI A, LAVAZZA A, REZZANI R et al. Mercuric chloride-induced alterations in stress protein distribution in rat kidney. Histol Histopathol 2004; 19:1209–1218. 119. NAVA M, ROMERO F, QUIROZ Y et al. Melatonin attenuates acute renal failure and oxidative stress induced by mercuric chloride in rats. Am J Physiol Renal Physiol 2000; 279: F910–F918. 120. WAALKES MP, LIU J, WARD JM et al. Mechanisms underlying arsenic carcinogenesis: hypersensitivity of mice exposed to inorganic arsenic during gestation. Toxicology 2004; 198:31–38. 121. BASU A, MAHATA J, GUPTA S et al. Genetic toxicology of a paradoxical human carcinogen, arsenic: a review. Mutat Res 2001; 488:171–194. 122. APOSHIAN HV. Enzymatic methylation of arsenic species and other new approaches to arsenic toxicity. Annu Rev Pharmacol Toxicol 1997; 37:397–419. 123. COHEN SM, ARNOLD LL, ELDAN M et al. Methylated arsenicals: the implications of metabolism and carcinogenicity studies in rodents to human risk assessment. Crit Rev Toxicol 2006; 36:99–133. 124. FLORA SJ. Arsenic-induced oxidative stress and its reversibility. Free Radic Biol Med 2011; 51:257–281. 125. ERCAL N, GURER-ORHAN H, AYKIN-BURNS N. Toxic metals and oxidative stress part I: mechanisms involved in metalinduced oxidative damage. Curr Top Med Chem 2001; 1:529–539. 126. SHI H, SHI X, LIU KJ. Oxidative mechanism of arsenic toxicity and carcinogenesis. Mol Cell Biochem 2004; 255: 67–78. 127. GURR JR, YIH LH, SAMIKKANNU T et al. Nitric oxide production by arsenite. Mutat Res 2003; 533:173–182. 128. NORDENSON I, BECKMAN L. Is the genotoxic effect of arsenic mediated by oxygen free radicals? Hum Hered 1991; 41: 71–73. 129. WANG TS, HUANG H. Active oxygen species are involved in the induction of micronuclei by arsenite in XRS-5 cells. Mutagenesis 1994; 9:253–257. 130. LIU S, WANG F, YAN L et al. Oxidative stress and MAPK involved into ATF2 expression in immortalized human urothelial cells treated by arsenic. Arch Toxicol 2013; 87: 981–989. 131. WANG H, XI S, XU Y et al. Sodium arsenite induces cyclooxygenase-2 expression in human uroepithelial cells through MAPK pathway activation and reactive oxygen species induction. Toxicol In Vitro 2013; 27:1043–1048. 132. KITCHIN KT, CONOLLY R. Arsenic-induced carcinogenesis– oxidative stress as a possible mode of action and future research needs for more biologically based risk assessment. Chem Res Toxicol 2010; 23:327–335. 133. ROSSMAN TG. Mechanism of arsenic carcinogenesis: an integrated approach. Mutat Res 2003; 533:37–65. 134. NESNOW S, ROOP BC, LAMBERT G et al. DNA damage induced by methylated trivalent arsenicals is mediated by reactive oxygen species. Chem Res Toxicol 2002; 15: 1627–1634. 135. WEI M, ARNOLD L, CANO M et al. Effects of co-administration of antioxidants and arsenicals on the rat urinary bladder epithelium. Toxicol Sci 2005; 83:237–245.

22

136. PAL S, CHATTERJEE AK. Prospective protective role of melatonin against arsenic-induced metabolic toxicity in Wistar rats. Toxicology 2005; 208:25–33. 137. PAL S, CHATTERJEE AK. Possible beneficial effects of melatonin supplementation on arsenic-induced oxidative stress in Wistar rats. Drug Chem Toxicol 2006; 29: 423–433. 138. LIN AM, FANG SF, CHAO PL et al. Melatonin attenuates arsenite-induced apoptosis in rat brain: involvement of mitochondrial and endoplasmic reticulum pathways and aggregation of alpha-synuclein. J Pineal Res 2007; 43: 163–171. 139. LIN AM, FENG SF, CHAO PL et al. Melatonin inhibits arsenite-induced peripheral neurotoxicity. J Pineal Res 2009; 46:64–70. 140. UYGUR R, AKTAS C, CAGLAR V et al. Protective effects of melatonin against arsenic-induced apoptosis and oxidative stress in rat testes. Toxicol Ind Health 2013. doi: 10. 1177/0748233713512891. 141. FAN SF, CHAO PL, LIN AM. Arsenite induces oxidative injury in rat brain: synergistic effect of iron. Ann N Y Acad Sci 2010; 1199:27–35. 142. PANT HH, RAO MV. Evaluation of in vitro anti-genotoxic potential of melatonin against arsenic and fluoride in human blood cultures. Ecotoxicol Environ Saf 2010; 73:1333–1337. 143. OBERLEY TD, FRIEDMAN AL, MOSER R et al. Effects of lead administration on developing rat kidney. II. Functional, morphologic, and immunohistochemical studies. Toxicol Appl Pharmacol 1995; 131:94–107. 144. Van Den HEUVEL RL, LEPPENS H, SCHOETERS GE. Use of in vitro assays to assess hematotoxic effects of environmental compounds. Cell Biol Toxicol 2001; 17:107–116. 145. JOHNSON FM. The genetic effects of environmental lead. Mutat Res 1998; 410:123–140. 146. FINKELSTEIN Y, MARKOWITZ ME, ROSEN JF. Low-level lead-induced neurotoxicity in children: an update on central nervous system effects. Brain Res Brain Res Rev 1998; 27:168–176. 147. BEYERSMANN D, HARTWIG A. Carcinogenic metal compounds: recent insight into molecular and cellular mechanisms. Arch Toxicol 2008; 82:493–512. 148. RIO B, FROQUET R, PARENT-MASSIN D. In vitro effect of lead acetate on human erythropoietic progenitors. Cell Biol Toxicol 2001; 17:41–50. 149. PATRICK L. Lead toxicity part II: the role of free radical damage and the use of antioxidants in the pathology and treatment of lead toxicity. Altern Med Rev 2006; 11: 114–127. 150. REITER RJ, TAN DX, CABRERA J et al. The oxidant/antioxidant network: role of melatonin. Biol Signals Recept 1999; 8:56–63. 151. SURESH C, DENNIS AO, HEINZ J et al. Melatonin protection against lead-induced changes in human neuroblastoma cell cultures. Int J Toxicol 2006; 25:459–464. 152. SLIWINSKI T, ROZEJ W, MORAWIEC-BAJDA A et al. Protective action of melatonin against oxidative DNA damage: chemical inactivation versus base-excision repair. Mutat Res 2007; 634:220–227. 153. MARTINEZ-ALFARO M, HERNANDEZ-CORTES D, WROBEL K et al. Effect of melatonin administration on DNA damage and repair responses in lymphocytes of rats subchronically exposed to lead. Mutat Res 2012; 742:37–42.

Melatonin for protection against metals 154. EL-SOKKARY GH, ABDEL-RAHMAN GH, KAMEL ES. Melatonin protects against lead-induced hepatic and renal toxicity in male rats. Toxicology 2005; 213:25–33. 155. KAWAHARA M. Effects of aluminum on the nervous system and its possible link with neurodegenerative diseases. J Alzheimers Dis 2005; 8:171–182. € 156. RIIHIMAKI V, AITIO A. Occupational exposure to aluminum and its biomonitoring in perspective. Crit Rev Toxicol 2012; 42:827–853. € 157. MEYER-BARON M, SCHAPER M, KNAPP G et al. Occupational aluminum exposure: evidence in support of its neurobehavioral impact. Neurotoxicology 2007; 28:1068–1078. 158. ZATTA P, LUCCHINI R, Van RENSBURG SJ et al. The role of metals in neurodegenerative processes: aluminum, manganese, and zinc. Brain Res Bull 2003; 62:15–28. € 159. IREGREN A, SJOGREN B, GUSTAFSSON K et al. Effects on the nervous system in different groups of workers exposed to aluminium. Occup Environ Med 2001; 58:453–460. 160. VERSTRAETEN SV, AIMO L, OTEIZA PI. Aluminium and lead: molecular mechanisms of brain toxicity. Arch Toxicol 2008; 82:789–802. 161. GANROT P. Metabolism and possible health effects of aluminum. Environ Health Persp 1986; 65:363–441. 162. GONCß ALVES PP, SILVA VS. Does neurotransmission impairment accompany aluminium neurotoxicity? J Inorg Biochem 2007; 101:1291–1338. 163. KAWAHARA M, MURAMOTO K, KOBAYASHI K et al. Functional and morphological changes in cultured neurons of rat cerebral cortex induced by long-term application of aluminum. Biochem Biophys Res Commun 1992; 189: 1317–1322. 164. SHCHERBATYKH I, CARPENTER DO. The role of metals in the etiology of Alzheimer’s disease. J Alzheimers Dis 2007; 11:191–205. 165. TOMLJENOVIC L. Aluminum and Alzheimer’s disease: after a century of controversy, is there a plausible link? J Alzheimers Dis 2011; 23:567–598. 166. ESPARZA J, GOMEZ M, ROMEU M et al. Aluminum-induced pro-oxidant effects in rats: protective role of exogenous melatonin. J Pineal Res 2003; 35:32–39. 167. ESPARZA JL, GOMEZ M, ROSA NOGUES M et al. Melatonin reduces oxidative stress and increases gene expression in the cerebral cortex and cerebellum of aluminum-exposed rats. J Pineal Res 2005; 39:129–136.   MR et al. Pro-oxidant 168. GOMEZ M, ESPARZA JL, NOGUES activity of aluminum in the rat hippocampus: gene expression of antioxidant enzymes after melatonin administration. Free Radic Biol Med 2005; 38:104–111.  MR et al. Oxidative stress 169. GARCIA T, ESPARZA JL, NOGUES status and RNA expression in hippocampus of an animal model of Alzheimer’s disease after chronic exposure to aluminum. Hippocampus 2010; 20:218–225. 170. GARCIA T, ESPARZA JL, GIRALT M et al. Protective role of melatonin on oxidative stress status and RNA expression in cerebral cortex and cerebellum of abpp transgenic mice after chronic exposure to aluminum. Biol Trace Elem Res 2010; 135:220–232. 171. ABD-ELGHAFFAR S, EL-SOKKARY GH, SHARKAWY AA. Aluminum-induced neurotoxicity and oxidative damage in rabbits: protective effect of melatonin. Neuro Endocrinol Lett 2005; 26:609–616. 172. PRIYANKA S, MAHITHA B, SRAVANTHI P et al. Melatonin prevents aluminum-induced oxidative damage and cytotoxicity

173.

174.

175.

176.

177.

178.

179.

180.

181.

182. 183.

184.

185.

186.

187.

188.

189.

in the medulla oblongata of mouse brain. J Pharm Res 2011; 4:683–686.  -PLANO S, GARCIA JJ, MARTINEZ-BALLARIN E et al. MILLAN Melatonin and pinoline prevent aluminium-induced lipid peroxidation in rat synaptosomes. J Trace Elem Med Biol 2003; 17:39–44. VanRENSBURG SJ, DANIELS WM, POTOCNIK FC et al. A new model for the pathophysiology of Alzheimer’s disease. Aluminium toxicity is exacerbated by hydrogen peroxide and attenuated by an amyloid protein fragment and melatonin. S Afr Med J 1997; 87:1111–1115. DANIELS W, RENSBURG SV, ZYL JV et al. Melatonin prevents b-amyloid-induced lipid peroxidation. J Pineal Res 1998; 24:78–82.   ALBENDEA CD, GOMEZ -TRULLEN EM, FUENTES-BROTO L et al. Melatonin reduces lipid and protein oxidative damage in synaptosomes due to aluminium. J Trace Elem Med Biol 2007; 21:261–268. GUTTERIDGE J, QUINLAN GJ, CLARK I et al. Aluminium salts accelerate peroxidation of membrane lipids stimulated by iron salts. Bichim Biophys Acta 1985; 835:441–447. GARCÍ A JNJ, REITER RJ, GUERRERO JM et al. Melatonin prevents changes in microsomal membrane fluidity during induced lipid peroxidation. FEBS Lett 1997; 408: 297–300. GARCIA J, REITER R, ORTIZ G et al. Melatonin enhances tamoxifen’s ability to prevent the reduction in microsomal membrane fluidity induced by lipid peroxidation. J Membr Biol 1998; 162:59–65. GARCIA J, REITER R, PIE J et al. Role of pinoline and melatonin in stabilizing hepatic microsomal membranes against oxidative stress. J Bioenerg Biomembr 1999; 31:609–616. ALFREY AC, HEGG A, CRASWELL P. Metabolism and toxicity of aluminum in renal failure. Am J Clin Nutr 1980; 33:1509–1516. ALFREY A. Aluminum metabolism in uremia. Neurotoxicology 1980; 1:43–53. S ß AHIN G, VAROL I, TEMIZER A et al. Determination of aluminum levels in the kidney, liver, and brain of mice treated with aluminum hydroxide. Biol Trace Elem Res 1994; 41:129–135.  MAHIEU S, MILLEN N, GONZALEZ M et al. Alterations of the renal function and oxidative stress in renal tissue from rats chronically treated with aluminium during the initial phase of hepatic regeneration. J Inorg Biochem 2005; 99:1858–1864.  MAHIEU S, CONTINI MDC, GONZALEZ M et al. Melatonin reduces oxidative damage induced by aluminium in rat kidney. Toxicol Lett 2009; 190:9–15.  DEL CARMEN CONTINI M, MILLEN N, GONZALEZ M et al. Melatonin prevents oxidative stress in ovariectomized rats treated with aluminium. Biol Trace Elem Res 2011; 144:924–943. STANA LG, MUSELIN F, CRETESCU I et al. The influence of sempervivum tectorum and melatonin administration on erythrocyte catalase in rats exposed to aluminium sulphate. Animal Sci Biotechnol 2012; 45:133–136. ABDEL-WAHAB WM. AlCl3-induced toxicity and oxidative stress in liver of male rats: protection by melatonin. Life Sci J 2012; 9:1173–1182. LACK B, DAYA S, NYOKONG T. Interaction of serotonin and melatonin with sodium, potassium, calcium, lithium and aluminium. J Pineal Res 2001; 31:102–108.

23

Romero et al. 190. GARCIA T, RIBES D, COLOMINA MT et al. Evaluation of the protective role of melatonin on the behavioral effects of aluminum in a mouse model of Alzheimer’s disease. Toxicology 2009; 265:49–55. 191. ISHIKAWA Y, NAKAGAWA K, SATOH Y et al. Characteristics of chromate workers’ cancers, chromium lung deposition and precancerous bronchial lesions: an autopsy study. Br J Cancer 1994; 70:160–166. 192. FRANCHINI I, MAGNANI F, MUTTI A. Mortality experience among chromeplating workers. Initial findings. Scand J Work Environ Health 1983; 9:247–252. 193. BARUTHIO F. Toxic effects of chromium and its compounds. Biol Trace Elem Res 1992; 32:145–153. 194. BRIGHT P, BURGE PS, O’HICKEY SP et al. Occupational asthma due to chrome and nickel electroplating. Thorax 1997; 52:28–32. 195. BUTTNER B, BEYERSMANN D. Modification of the erythrocyte anion carrier by chromate. Xenobiotica 1985; 15: 735–741. 196. STANDEVEN AM, WETTERHAHN KE. Ascorbate is the principal reductant of chromium(VI) in rat lung ultrafiltrates and cytosols, and mediates chromium-DNA binding in vitro. Carcinogenesis 1992; 13:1319–1324. 197. SHI XL, DALAL NS. NADPH-dependent flavoenzymes catalyze one electron reduction of metal ions and molecular oxygen and generate hydroxyl radicals. FEBS Lett 1990; 276:189–191. 198. AIYAR J, BERKOVITS HJ, FLOYD RA et al. Reaction of chromium (VI) with hydrogen peroxide in the presence of glutathione: reactive intermediates and resulting DNA damage. Chem Res Toxicol 1990; 3:595–603. 199. MYERS JM, MYERS CR. The effects of hexavalent chromium on thioredoxin reductase and peroxiredoxins in human bronchial epithelial cells. Free Radic Biol Med 2009; 47:1477–1485. 200. MYERS JM, ANTHOLINE WE, MYERS CR. The intracellular redox stress caused by hexavalent chromium is selective for proteins that have key roles in cell survival and thiol redox control. Toxicology 2011; 281:37–47. 201. TAN DX, MANCHESTER LC, TERRON MP et al. One molecule, many derivatives: a never-ending interaction of melatonin with reactive oxygen and nitrogen species? J Pineal Res 2007; 42:28–42. 202. GALANO A, TAN DX, REITER RJ. Cyclic 3-hydroxymelatonin, a key metabolite enhancing the peroxyl radical scavenging activity of melatonin. RSC Adv 2014; 4: 5220–5227. 203. QI W, REITER RJ, TAN DX et al. Chromium(III)-induced 8-hydroxydeoxyguanosine in DNA and its reduction by antioxidants: comparative effects of melatonin, ascorbate, and vitamin E. Environ Health Perspect 2000; 108:399–402. 204. QI W, REITER RJ, TAN DX et al. Increased levels of oxidatively damaged DNA induced by chromium(III) and H2O2: protection by melatonin and related molecules. J Pineal Res 2000; 29:54–61. 205. LOPEZ-BURILLO S, TAN DX, MAYO JC et al. Melatonin, xanthurenic acid, resveratrol, EGCG, vitamin C and alphalipoic acid differentially reduce oxidative DNA damage induced by Fenton reagents: a study of their individual and synergistic actions. J Pineal Res 2003; 34:269–277. 206. SUSA N, UENO S, FURUKAWA Y et al. Potent protective effect of melatonin on chromium(VI)-induced DNA singlestrand breaks, cytotoxicity, and lipid peroxidation in

24

207. 208.

209. 210.

211.

212. 213. 214.

215.

216.

217. 218.

219. 220.

221.

222.

223.

224.

225. 226. 227.

228.

primary cultures of rat hepatocytes. Toxicol Appl Pharmacol 1997; 144:377–384. ANDREWS NC. Disorders of iron metabolism. N Engl J Med 1999; 341:1986–1995. KOPPENOL WH. Free radical damage and its control. In: Chemistry of Iron and Copper in Radical Reactions. RICE-EVANS CA, BURDON RH, ed., Elsevier Science B. V., Burdon, Amsterdam, 1994; pp. 3–24. FRAGA CG, OTEIZA PI. Iron toxicity and antioxidant nutrients. Toxicology 2002; 180:23–32. FAHN S, COHEN G. The oxidant stress hypothesis in Parkinson’s disease: evidence supporting it. Ann Neurol 1992; 32:804–812. SMITH MA, HARRIS PL, SAYRE LM et al. Iron accumulation in Alzheimer disease is a source of redox-generated free radicals. Proc Natl Acad Sci USA 1997; 94:9866–9868. KEHRER JP. The Haber–Weiss reaction and mechanisms of toxicity. Toxicology 2000; 149:43–50. GUTTERIDGE JM, HALLIWELL B. Iron toxicity and oxygen radicals. Baillieres Clin Haematol 1989; 2:195–256. LIOCHEV SI. The mechanism of “Fenton-like” reactions and their importance for biological systems. A biologist’s view. Met Ions Biol Syst 1999; 36:1–39. HABER F, WEISS J. The catalytic decomposition of hydrogen peroxide by iron salts. Proc R Soc Lond A Math Phys Sci 1934; 147:332–351. WISEMAN H, HALLIWELL B. Damage to DNA by reactive oxygen and nitrogen species: role in inflammatory disease and progression to cancer. Biochem J 1996; 313(Pt 1): 17–29. COOPER CE. Nitric oxide and iron proteins. Biochim Biophys Acta 1999; 1411:290–309. BROWN GC. Regulation of mitochondrial respiration by nitric oxide inhibition of cytochrome c oxidase. Biochim Biophys Acta 2001; 1504:46–57. AKTAN F. iNOS-mediated nitric oxide production and its regulation. Life Sci 2004; 75:639–654. SOLTES L, KOGAN G. Catabolism of hyaluronan: involvement of transition metals. Interdiscip Toxicol 2009; 2: 229–238. BUCHER JR, TIEN M, AUST SD. The requirement for ferric in the initiation of lipid peroxidation by chelated ferrous iron. Biochem Biophys Res Commun 1983; 111:777–784. MINOTTI G, AUST SD. The requirement for iron (III) in the initiation of lipid peroxidation by iron (II) and hydrogen peroxide. J Biol Chem 1987; 262:1098–1104. LOVEJOY DB, RICHARDSON DR. Iron chelators as antineoplastic agents: current developments and promise of the PIH class of chelators. Curr Med Chem 2003; 10: 1035–1049. RICHARDSON D, PONKA P, BAKER E. The effect of the iron (III) chelator, desferrioxamine, on iron and transferrin uptake by the human malignant melanoma cell. Cancer Res 1994; 54:685–689. LIU ZD, HIDER RC. Design of clinically useful iron(III)selective chelators. Med Res Rev 2002; 22:26–64. LIU ZD, HIDER RC. Design of iron chelators with therapeutic application. Coord Chem Rev 2002; 232:151–171. NUNES A, MARQUES S, QUINTANOVA C et al. Multifunctional iron-chelators with protective roles against neurodegenerative diseases. Dalton Trans 2013; 42:6058–6073. Van Der SCHYF CJ, GAL S, GELDENHUYS WJ et al. Multifunctional neuroprotective drugs targeting monoamine

Melatonin for protection against metals

229.

230.

231.

232.

233.

234.

235.

236. 237.

238.

239.

240.

241.

242. 243.

244.

245.

246.

247.

oxidase inhibition, iron chelation, adenosine receptors, and cholinergic and glutamatergic action for neurodegenerative diseases. Expert Opin Investig Drugs 2006; 15:873–886. WEINREB O, MANDEL S, BAR-AM O et al. Iron-chelating backbone coupled with monoamine oxidase inhibitory moiety as novel pluripotential therapeutic agents for Alzheimer’s disease: a tribute to Moussa Youdim. J Neural Transm 2011; 118:479–492. RICHARDSON DR. Novel chelators for central nervous system disorders that involve alterations in the metabolism of iron and other metal ions. Ann N Y Acad Sci 2006; 1012:326–341. WEINREB O, MANDEL S, YOUDIM MB et al. Targeting dysregulation of brain iron homeostasis in Parkinson’s disease by iron chelators. Free Radic Biol Med 2013; 62:52–64. MARTIN M, MACIAS M, ESCAMES G et al. Melatonin but not vitamins C and E maintains glutathione homeostasis in t-butyl hydroperoxide-induced mitochondrial oxidative stress. FASEB J 2000; 14:1677–1679. BARLOW-WALDEN L, REITER R, ABE M et al. Melatonin stimulates brain glutathione peroxidase activity. Neurochem Int 1995; 26:497–502. PABLOS MI, REITER RJ, ORTIZ GG et al. Rhythms of glutathione peroxidase and glutathione reductase in brain of chick and their inhibition by light. Neurochem Int 1998; 32:69–75. KOTLER M, RODRIGUEZ C, SAINZ RM et al. Melatonin increases gene expression for antioxidant enzymes in rat brain cortex. J Pineal Res 1998; 24:83–89. CHANCE B, SIES H, BOVERIS A. Hydroperoxide metabolism in mammalian organs. Physiol Rev 1979; 59:527–605. TAN D, CHEN L, POEGGELER B et al. Melatonin: a potent, endogenous hydroxyl radical scavenger. Endocr J 1993; 1:57–60. PIERI C, MARRA M, MORONI F et al. Melatonin: a peroxyl radical scavenger more effective than vitamin E. Life Sci 1994; 55:PL271–PL276. ~ -CASTROVIEJO D, TAN DX et al. Free radREITER RJ, ACUNA ical-mediated molecular damage. Ann N Y Acad Sci 2001; 939:200–215. REITER RJ. Functional aspects of the pineal hormone melatonin in combating cell and tissue damage induced by free radicals. Eur J Endocrinol 1996; 134:412–420. TAN DX, MANCHESTER LC, REITER RJ et al. Significance of melatonin in antioxidative defense system: reactions and products. Biol Signals Recept 2000; 9:137–159. MARKESBERY WR. Oxidative stress hypothesis in Alzheimer’s disease. Free Radic Biol Med 1997; 23:134–147. DEXTER D, WELLS F, LEE A et al. Increased nigral iron content and alterations in other metal ions occurring in brain in Parkinson’s disease. J Neurochem 1989; 52:1830–1836. WILLMORE LJ, SYPERT GW, MUNSON JB. Recurrent seizures induced by cortical iron injection: a model of posttraumatic epilepsy. Ann Neurol 1978; 4:329–336. KABUTO H, YOKOI I, OGAWA N. Melatonin inhibits ironinduced epileptic discharges in rats by suppressing peroxidation. Epilepsia 1998; 39:237–243. HAYTER CL, BISHOP GM, ROBINSON SR. Pharmacological but not physiological concentrations of melatonin reduce iron-induced neuronal death in rat cerebral cortex. Neurosci Lett 2004; 362:182–184. SOFIC E, PAULUS W, JELLINGER K et al. Selective increase of iron in substantia nigra zona compacta of parkinsonian brains. J Neurochem 1991; 56:978–982.

248. LIN AM, HO LT. Melatonin suppresses iron-induced neurodegeneration in rat brain. Free Radic Biol Med 2000; 28:904–911. 249. CHEN KB, LIN AM, CHIU TH. Oxidative injury to the locus coeruleus of rat brain: neuroprotection by melatonin. J Pineal Res 2003; 35:109–117. 250. EHMANN W, MARKESBERY W, ALAUDDIN M et al. Brain trace elements in Alzheimer’s disease. Neurotoxicology 1986; 7:195. 251. GOODMAN L. Alzheimer’s disease: a clinico-pathologic analysis of twenty-three cases with a theory on pathogenesis. J Nerv Ment Dis 1953; 118:97–130. 252. THOMPSON C, MARKESBERY W, EHMANN W et al. Regional brain trace-element studies in Alzheimer’s disease. Neurotoxicology 1988; 9:1–7. 253. OZCANKAYA R, DELIBAS N. Malondialdehyde, superoxide dismutase, melatonin, iron, copper, and zinc blood concentrations in patients with Alzheimer disease: cross-sectional study. Croat Med J 2002; 43:28–32. 254. QI W, REITER RJ, TAN DX et al. Melatonin prevents deltaaminolevulinic acid-induced oxidative DNA damage in the presence of Fe2+. Mol Cell Biochem 2001; 218:87–92. 255. TAN D-X, MANCHESTER LC, REITER RJ et al. Melatonin directly scavenges hydrogen peroxide: a potentially new metabolic pathway of melatonin biotransformation. Free Radic Biol Med 2000; 29:1177–1185. 256. MENENDEZ-PELAEZ A, POEGGELER B, REITER RJ et al. Nuclear localization of melatonin in different mammalian tissues: immunocytochemical and radioimmunoassay evidence. J Cell Biochem 1993; 53:373–382. 257. MENENDEZ-PELAEZ A, REITER RJ. Distribution of melatonin in mammalian tissues: the relative importance of nuclear versus cytosolic localization. J Pineal Res 1993; 15:59–69. 258. MUNOZ-CASTANEDA JR, TUNEZ I, HERENCIA C et al. Melatonin exerts a more potent effect than S-adenosyl-l-methionine against iron metabolism disturbances, oxidative stress and tissue injury induced by obstructive jaundice in rats. Chem Biol Interact 2008; 174:79–87. 259. MILCZAREK R, HALLMANN A, SOKOOWSKA E et al. Melatonin enhances antioxidant action of a-tocopherol and ascorbate against NADPH-and iron-dependent lipid peroxidation in human placental mitochondria. J Pineal Res 2010; 49:149–155. 260. DEMANT EJ. Transfer of ferritin-bound iron to adriamycin. FEBS Lett 1984; 176:97–100. 261. OTHMAN AI, EL-MISSIRY MA, AMER MA et al. Melatonin controls oxidative stress and modulates iron, ferritin, and transferrin levels in adriamycin treated rats. Life Sci 2008; 83:563–568. 262. UAUY R, OLIVARES M, GONZALEZ M. Essentiality of copper in humans. Am J Clin Nutr 1998; 67:952S–959S. ~ 263. ARREDONDO M, NU NEZ MT. Iron and copper metabolism. Mol Aspects Med 2005; 26:313–327. 264. O’HALLORAN TV, CULOTTA VC. Metallochaperones, an intracellular shuttle service for metal ions. J Biol Chem 2000; 275:25057–25060. 265. MERCER JFB. The molecular basis of copper-transport diseases. Trends Mol Med 2001; 7:64–69. 266. CRICHTON RR, DEXTER DT, WARD RJ. Metal based neurodegenerative diseases—From molecular mechanisms to therapeutic strategies. Coord Chem Rev 2008; 252: 1189–1199.

25

Romero et al. 267. LIOCHEV SI, FRIDOVICH I. The Haber–Weiss cycle – 70 years later: an alternative view. Redox Rep 2002; 7:55–57; author reply 59–60. 268. SUTTON HC, WINTERBOURN CC. On the participation of higher oxidation states of iron and copper in Fenton reactions. Free Radic Biol Med 1989; 6:53–60. 269. ARUOMA OI, HALLIWELL B, GAJEWSKI E et al. Copper-iondependent damage to the bases in DNA in the presence of hydrogen peroxide. Biochem J 1991; 273(Pt 3):601–604. 270. SPEISKY H, GOMEZ M, BURGOS-BRAVO F et al. Generation of superoxide radicals by copper-glutathione complexes: redox-consequences associated with their interaction with reduced glutathione. Bioorg Med Chem 2009; 17: 1803–1810. 271. LINDER MC, HAZEGH-AZAM M. Copper biochemistry and molecular biology. Am J Clin Nutr 1996; 63:797S–811S. 272. URIU-ADAMS JY, KEEN CL. Copper, oxidative stress, and human health. Mol Aspects Med 2005; 26:268–298. 273. BREZOVA V, VALKO M, BREZA M et al. Role of radicals and singlet oxygen in photoactivated dna cleavage by the anticancer drug camptothecin: an electron paramagnetic resonance study. J Phys Chem B 2003; 107:2415–2425.  274. LETELIER ME, SANCHEZ -JOFRE S, PEREDO-SILVA L et al. Mechanisms underlying iron and copper ions toxicity in biological systems: pro-oxidant activity and protein-binding effects. Chem Biol Interact 2010; 188:220–227. 275. RAE T, SCHMIDT P, PUFAHL R et al. Undetectable intracellular free copper: the requirement of a copper chaperone for superoxide dismutase. Science 1999; 284:805–808. 276. BARBOUTI A, DOULIAS P-T, ZHU B-Z et al. Intracellular iron, but not copper, plays a critical role in hydrogen peroxide-induced DNA Damage. Free Radic Biol Med 2001; 31:490–498. 277. BREWER GJ. The risks of free copper in the body and the development of useful anticopper drugs. Curr Opin Clin Nutr Metab Care 2008; 11:727–732.  278. LETELIER ME, LEPE AM, FAUNDEZ M et al. Possible mechanisms underlying copper-induced damage in biological membranes leading to cellular toxicity. Chem Biol Interact 2005; 151:71–82. 279. GAGGELLI E, KOZLOWSKI H, VALENSIN D et al. Copper homeostasis and neurodegenerative disorders (Alzheimer’s, prion, and Parkinson’s diseases and amyotrophic lateral sclerosis). Chem Rev 2012; 2010:1995–2044. 280. DONNELLY PS, XIAO Z, WEDD AG. Copper and Alzheimer’s disease. Curr Opin Chem Biol 2007; 11:128–133. 281. BARNHAM KJ, CAPPAI R, BEYREUTHER K et al. Delineating common molecular mechanisms in Alzheimer’s and prion diseases. Trends Biochem Sci 2006; 31:465–472. 282. FOX JH, KAMA JA, LIEBERMAN G et al. Mechanisms of copper ion mediated Huntington’s disease progression. PLoS ONE 2007; 2:e334. 283. VALENTINE JS, DOUCETTE PA, ZITTIN POTTER S. Copper-zinc superoxide dismutase and amyotrophic lateral sclerosis. Annu Rev Biochem 2005; 74:563–593. 284. ROBERTS RA, SMITH RA, SAFE S et al. Toxicological and pathophysiological roles of reactive oxygen and nitrogen species. Toxicology 2010; 276:85–94. 285. GUPTE A, MUMPER RJ. Elevated copper and oxidative stress in cancer cells as a target for cancer treatment. Cancer Treat Rev 2009; 35:32–46. 286. HUNG YH, BUSH AI, CHERNY RA. Copper in the brain and Alzheimer’s disease. J Biol Inorg Chem 2010; 15:61–76.

26

287. COOPER GJ, PHILLIPS AR, CHOONG SY et al. Regeneration of the heart in diabetes by selective copper chelation. Diabetes 2004; 53:2501–2508. 288. HAIDARI M, JAVADI E, KADKHODAEE M et al. Enhanced susceptibility to oxidation and diminished vitamin E content of LDL from patients with stable coronary artery disease. Clin Chem 2001; 47:1234–1240. 289. CUNNINGHAM J, LEFFELL M, MEARKLE P et al. Elevated plasma ceruloplasmin in insulin-dependent diabetes mellitus: evidence for increased oxidative stress as a variable complication. Metabolism 1995; 44:996–999. 290. PARMAR P, LIMSON J, NYOKONG T et al. Melatonin protects against copper-mediated free radical damage. J Pineal Res 2002; 32:237–242. 291. HANDY RD. Chronic effects of copper exposure versus endocrine toxicity: two sides of the same toxicological process? Comp Biochem Physiol A Mol Integr Physiol 2003; 135:25–38. 292. PARMAR P, DAYA S. The effect of copper on (3H)-tryptophan metabolism in organ cultures of rat pineal glands. Metab Brain Dis 2001; 16:199–205. 293. WAKATSUKI A, OKATANI Y, IKENOUE N et al. Melatonin inhibits oxidative modification of low-density lipoprotein particles in normolipidemic post-menopausal women. J Pineal Res 2000; 28:136–142. 294. PIERI C, MARRA M, GASPAR R et al. Melatonin protects LDL from oxidation but does not prevent the apolipoprotein derivatization. Biochem Biophys Res Commun 1996; 222:256–260. 295. KELLY MR, LOO G. Melatonin inhibits oxidative modification of human low-density lipoprotein. J Pineal Res 1997; 22:203–209. 296. WALTERS-LAPORTE E, FURMAN C, FOUQUET S et al. A high concentration of melatonin inhibits in vitro LDL peroxidation but not oxidized LDL toxicity toward cultured endothelial cells. J Cardiovasc Pharmacol 1998; 32:582–592. 297. MAYO JC, TAN DX, SAINZ RM et al. Protection against oxidative protein damage induced by metal-catalyzed reaction or alkylperoxyl radicals: comparative effects of melatonin and other antioxidants. Biochim Biophys Acta 2003; 1620:139–150. 298. ZATTA P, TOGNON G, CARAMPIN P. Melatonin prevents free radical formation due to the interaction between beta-amyloid peptides and metal ions [Al(III), Zn(II), Cu(II), Mn (II), Fe(II)]. J Pineal Res 2003; 35:98–103. 299. KLAASSEN CD. Casarett & Doull’s Toxicology: The Basic Science of Poisons. McGraw-Hill, New York, 2001. 300. WRIGHT DA, WELBOURN P. Environmental Toxicology. In: Cambridge Environmental Chemistry Series, Vol. 11. PCG CAMPBELL. ed., Cambridge University Press, Cambridge, UK, 2002; pp. 304–306. 301. COOGAN TP, LATTA DM, SNOW ET et al. Toxicity and carcinogenicity of nickel compounds. Crit Rev Toxicol 1989; 19:341–384. 302. BARCELOUX DG. Nickel. J Toxicol Clin Toxicol 1999; 37:239–258. 303. ZAMPONI GW, BOURINET E, SNUTCH TP. Nickel block of a family of neuronal calcium channels: subtype- and subunitdependent action at multiple sites. J Membr Biol 1996; 151:77–90. 304. GOEBELER M, ROTH J, MEINARDUS-HAGER G et al. The Contact Allergens Nickel Chloride and Cobalt Chloride

Melatonin for protection against metals

305.

306.

307.

308.

309.

310.

311.

312. 313. 314.

315. 316.

317.

318.

319.

320.

321.

322.

Directly Induce Expression of Endothelial Adhesion Molecules. Behring Inst Mitt 1993; 92:191–201. VALKO M, RHODES CJ, MONCOL J et al. Free radicals, metals and antioxidants in oxidative stress-induced cancer. Chem Biol Interact 2006; 160:1–40. BASKETTER DA, BRIATICO-VANGOSA G, KAESTNER W et al. Nickel, cobalt and chromium in consumer products: a role in allergic contact dermatitis? Contact Dermatitis 1993; 28:15–25. CHEN CY, SU YJ, WU PF et al. Nickel-induced plasma lipid peroxidation and effect of antioxidants in human blood: involvement hydroxyl radical formation and depletion of alpha-tocopherol. J Toxicol Environ Health Part A 2002; 65:843–852. SALNIKOW K, AN WG, MELILLO G et al. Nickel-induced transformation shifts the balance between HIF-1 and p53 transcription factors. Carcinogenesis 1999; 20:1819–1823. GOLDBERG MA, DUNNING SP, BUNN HF. Regulation of the erythropoietin gene: evidence that the oxygen sensor is a heme protein. Science 1988; 242:1412–1415. XU SC, HE MD, ZHONG M et al. Melatonin protects against nickel-induced neurotoxicity in vitro by reducing oxidative stress and maintaining mitochondrial function. J Pineal Res 2010; 49:86–94. XU SC, HE MD, LU YH et al. Nickel exposure induces oxidative damage to mitochondrial DNA in Neuro2a cells: the neuroprotective roles of melatonin. J Pineal Res 2011; 51:426–433. FREGERT S, RORSMAN H. Allergy to chromium, nickel and cobalt. Acta Derm Venereol 1966; 46:144–148. BARBORIK M. [Hematologic changes in workers employed in the production of hard metals]. Prac Lek 1967; 19:11–15. BAKER DH, PARR TM, AUGSPURGER NR. Oral iodine toxicity in chicks can be reversed by supplemental bromine. J Nutr 2003; 133:2309–2312. BEYERSMANN D, HARTWIG A. The genetic toxicology of cobalt. Toxicol Appl Pharmacol 1992; 115:137–145. SIROVER MA, LOEB LA. Infidelity of DNA synthesis in vitro: screening for potential metal mutagens or carcinogens. Science 1976; 194:1434–1436. HENGSTLER JG, BOLM-AUDORFF U, FALDUM A et al. Occupational exposure to heavy metals: DNA damage induction and DNA repair inhibition prove co-exposures to cadmium, cobalt and lead as more dangerous than hitherto expected. Carcinogenesis 2003; 24:63–73. BEARDEN LJ, COOKE FW. Growth inhibition of cultured fibroblasts by cobalt and nickel. J Biomed Mater Res 1980; 14:289–309. OLIVIERI G, HESS C, SAVASKAN E et al. Melatonin protects SHSY5Y neuroblastoma cells from cobalt-induced oxidative stress, neurotoxicity and increased beta-amyloid secretion. J Pineal Res 2001; 31:320–325. SCHOENFELD N, GREENBLAT Y, EPSTEIN O et al. Evidence relating the inhibitory effect of cobalt on the activity of delta-aminolevulinate synthase to the intracellular concentration of porphyrins. Biochem Pharmacol 1983; 32: 2333–2337. SCHOELLER DA, KOTAKE AN, LAMBERT GH et al. Comparison of the phenacetin and aminopyrine breath tests: effect of liver disease, inducers and cobaltous chloride. Hepatology 1985; 5:276–281. HASAN M, ALI S, ANWAR J. Cobalt-induced depletion of dopamine, norepinephrine & 5-hydroxytryptamine concen-

323.

324.

325.

326.

327.

328.

329.

330.

331. 332.

333.

334.

335.

336.

337.

338.

tration in different regions of the rat brain. Indian J Exp Biol 1980; 18:1051–1053. WANG G, HAZRA TK, MITRA S et al. Mitochondrial DNA damage and a hypoxic response are induced by CoCl(2) in rat neuronal PC12 cells. Nucleic Acids Res 2000; 28: 2135–2140. EPSTEIN AC, GLEADLE JM, MCNEILL LA et al. The C. elegans EGL-9 and mammalian homologs define a family of dioxygenases that regulate HIF by prolyl hydroxylation.Cell 2001; 107:43–54. AVRAMOVICH-TIROSH Y, AMIT T, BAR-AM O et al. Therapeutic targets and potential of the novel brain- permeable multifunctional iron chelator-monoamine oxidase inhibitor drug, M-30, for the treatment of Alzheimer’s disease. J Neurochem 2007; 100:490–502. WENSTRUP D, EHMANN WD, MARKESBERY WR. Trace element imbalances in isolated subcellular fractions of Alzheimer’s disease brains. Brain Res 1990; 533:125–131. Van Den BROEKE LT, GRASLUND A, NILSSON JL et al. Free radicals as potential mediators of metal-allergy: Ni2+- and Co2+-mediated free radical generation. Eur J Pharm Sci 1998; 6:279–286. URATA Y, HONMA S, GOTO S et al. Melatonin induces gamma-glutamylcysteine synthetase mediated by activator protein-1 in human vascular endothelial cells. Free Radic Biol Med 1999; 27:838–847. DAI M, CUI P, YU M et al. Melatonin modulates the expression of VEGF and HIF-1 alpha induced by CoCl2 in cultured cancer cells. J Pineal Res 2008; 44:121–126. CORTIZO AM, BRUZZONE L, MOLINUEVO S et al. A possible role of oxidative stress in the vanadium-induced cytotoxicity in the MC3T3E1 osteoblast and UMR106 osteosarcoma cell lines. Toxicology 2000; 147:89–99. KOSTOVA I. Titanium and vanadium complexes as anticancer agents. Anticancer Agents Med Chem 2009; 9:827–842. CHANDRA AK, GHOSH R, CHATTERJEE A et al. Effects of vanadate on male rat reproductive tract histology, oxidative stress markers and androgenic enzyme activities. J Inorg Biochem 2007; 101:944–956. CHEN F, DEMERS LM, VALLYATHAN V et al. Vanadate induction of NF-kappaB involves IkappaB kinase beta and SAPK/ERK kinase 1 in macrophages. J Biol Chem 1999; 274:20307–20312. GAO N, DING M, ZHENG JZ et al. Vanadate-induced expression of hypoxia-inducible factor 1 alpha and vascular endothelial growth factor through phosphatidylinositol 3-kinase/Akt pathway and reactive oxygen species. J Biol Chem 2002; 277:31963–31971. BAY BH, SIT KH, PARAMANANTHAM R et al. Hydroxyl free radicals generated by vanadyl[IV] induce cell blebbing in mitotic human Chang liver cells. Biometals 1997; 10:119–122. NAGI MN, MANSOUR MA, AL-SHABANAH OA et al. Melatonin inhibits the contractile effect of vanadate in the isolated pulmonary arterial rings of rats: possible role of hydrogen peroxide. J Biochem Mol Toxicol 2002; 16:273–278. KIERSZTAN A, WINIARSKA K, DROZAK J et al. Differential effects of vanadium, tungsten and molybdenum on inhibition of glucose formation in renal tubules and hepatocytes of control and diabetic rabbits: beneficial action of melatonin and N-acetylcysteine. Mol Cell Biochem 2004; 261:9–21. RETTIE AE, FISHER MB. Transformation Enzymes. Oxidative; Non-P450. In: Handbook of Drug Metabolism. WOOLF TF ed., Marcel Dekker, New York, 1999; pp. 131–151.

27

Romero et al. 339. WRIGHT RM, REPINE JE. The human molybdenum hydroxylase gene family: co-conspirators in metabolic free-radical generation and disease. Biochem Soc Trans 1997; 25:799–804. 340. CHAN S, GERSON B, SUBRAMANIAM S. The role of copper, molybdenum, selenium, and zinc in nutrition and health. Clin Lab Med 1998; 18:673–685.

28

341. FLORA SJ. Structural, chemical and biological aspects of antioxidants for strategies against metal and metalloid exposure. Oxid Med Cell Longev 2009; 2:191–206. 342. GALANO A, TAN DX, REITER RJ. Melatonin as a natural ally against oxidative stress: a physicochemical examination. J Pineal Res 2011; 51:1–16.

A review of metal-catalyzed molecular damage: protection by melatonin.

Metal exposure is associated with several toxic effects; herein, we review the toxicity mechanisms of cadmium, mercury, arsenic, lead, aluminum, chrom...
621KB Sizes 2 Downloads 3 Views