Academia.eduAcademia.edu
Review Availability, Toxicology and Medical Significance of Antimony Argyrios Periferakis 1,2,†, Ana Caruntu 3,4,*, Aristodemos-Theodoros Periferakis 1,†, Andreea-Elena Scheau 5, Ioana Anca Badarau 1, Constantin Caruntu 1,6 and Cristian Scheau 1,* Department of Physiology, The “Carol Davila” University of Medicine and Pharmacy, 050474 Bucharest, Romania; argyrios.periferakis0920@stud.umfcd.ro (A.P.); theodoros.periferakis0920@stud.umfcd.ro (A.-T.P.); anca.badarau@umfcd.ro (I.A.B.); costin.caruntu@gmail.com (C.C.) 2 Akadimia of Ancient Greek and Traditional Chinese Medicine, 16675 Athens, Greece 3 Department of Oral and Maxillofacial Surgery, The “Carol Davila” Central Military Emergency Hospital, 010825 Bucharest, Romania 4 Department of Oral and Maxillofacial Surgery, Faculty of Dental Medicine, “Titu Maiorescu” University, 031593 Bucharest, Romania 5 Department of Radiology and Medical Imaging, Fundeni Clinical Institute, 022328 Bucharest, Romania; andreea.ghergus@gmail.com 6 Department of Dermatology, Prof. N.C. Paulescu National Institute of Diabetes, Nutrition and Metabolic Diseases, 011233 Bucharest, Romania * Correspondence: ana.caruntu@gmail.com (A.C.); cristian.scheau@umfcd.ro (C.S.) † These authors contributed equally to this work. 1 Citation: Periferakis, A.; Caruntu, A.; Periferakis, A.-T.; Scheau, A.-E.; Badarau, I.A.; Caruntu, C.; Scheau, C. Availability, Toxicology and Medical Significance of Antimony. Int. J. Environ. Res. Public Health 2022, 19, 4669. https://doi.org/10.3390/ ijerph19084669 Academic Editor: José Antônio Menezes Filho Received: 1 March 2022 Abstract: Antimony has been known and used since ancient times, but its applications have increased significantly during the last two centuries. Aside from its few medical applications, it also has industrial applications, acting as a flame retardant and a catalyst. Geologically, native antimony is rare, and it is mostly found in sulfide ores. The main ore minerals of antimony are antimonite and jamesonite. The extensive mining and use of antimony have led to its introduction into the biosphere, where it can be hazardous, depending on its bioavailability and absorption. Detailed studies exist both from active and abandoned mining sites, and from urban settings, which document the environmental impact of antimony pollution and its impact on human physiology. Despite its evident and pronounced toxicity, it has also been used in some drugs, initially tartar emetics and subsequently antimonials. The latter are used to treat tropical diseases and their therapeutic potential for leishmaniasis means that they will not be soon phased out, despite the fact the antimonial resistance is beginning to be documented. The mechanisms by which antimony is introduced into human cells and subsequently excreted are still the subject of research; their elucidation will enable us to better understand antimony toxicity and, hopefully, to improve the nature and delivery method of antimonial drugs. Accepted: 10 April 2022 Published: 12 April 2022 Keywords: antimony; stibnite; toxicity; exposure; health impact; pathophysiology; resistance Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations. Copyright: © 2022 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license (https://creativecommons.org/licenses/by/4.0/). 1. Introduction Antimony (Sb), as an element, has been known since ancient times and has been used by many civilizations for different purposes. It is classified as a heavy metal, since it has a specific density of more than 5 gr/cm3 [1], and it has adverse effects on the health and physiology of living organisms [2]. Heavy metals such as antimony are released into the biosphere mostly via weathering and erosion, industrial and mining activities, and pest control agents [3]. Since antimony belongs to Group 15 of the periodic table, it is also referred to as a metalloid [4]. As a metal, antimony is not affected by humid air and pure water, but if melted by temperature, it ignites. It is known to react violently with elements of the halide group (F, Int. J. Environ. Res. Public Health 2022, 19, 4669. https://doi.org/10.3390/ijerph19084669 www.mdpi.com/journal/ijerph Int. J. Environ. Res. Public Health 2022, 19, 4669 2 of 30 Cl, Br, and I) thus forming trihalides. The detailed chemistry and properties of different phases of antimony are described in detail in [5]. Antimony has been known since antiquity in China and in Egypt, and there is proof that even the Chaldeans knew how to sequester it from other ores [6]. Accounts from Greek and Roman authors of the era refer to its main use as an eye ointment for its cosmetic and medicinal properties [5]. Around the 16th century, it was realized that antimony was useful in separating gold from silver. However, until the dawn of the 20th century, the demand for and use of antimony remained fairly low [7]. It would be during the First World War [8], and subsequently, during the Second World War, that demand for antimony increased, as it did for other materials such as oil [9] and emery [10–12]. After the 1950s, the dominant market of antimony became the plastics industry, which currently consumes over 60% of the produced antimony, both as a catalyst and a flame retardant. The pyrometallurgical, hydrometallurgical, electrometallurgical, and mineral processing processes currently employed at an industrial level are detailed in [13]. The usage of antimony derives from its particular properties: it is hard, brittle, and non-malleable, and this renders it unsuitable for being used in the same manner as other metals, such as Pb, Fe, etc. Instead, it is used in small amounts in alloys to enhance their strength and hardness, and it may be also used ornamentally. In the electronics industry, its use for diodes is increasing [13]. The basic artificial industrial compounds are antimony trioxide (Sb2O3), antimony pentoxide (Sb2O5), sodium antimonate (NaSbO3), antimony trisulfide (Sb2S3), antimony pentasulfide (Sb2S5), and antimony triacetate [Sb(CH3COOH)3]. They are mostly used as flame retardants and catalysts (Table 1). Table 1. Industrial uses of antimony compounds. Antimony Compound Chemical Formula Antimony trioxide Sb2O3 Antimony pentoxide Sb2O5 Sodium antimonate NaSbO3 Antimony trisulfide Antimony pentasulfide Antimony triacetate Sb2S3 Sb2S5 Sb(CH3COOH)3 Uses Flame retardant in plastics, textiles and rubber; catalyst for PET production Flame retardant Flame retardant; decolorizing and refining agent for optical glass Photoconductors, brake linings, fireworks Vulcanizing agent Catalyst in the production of polyesters Recently, antimony was included in the critical raw materials (CRM) list. These are materials that are characterized by increased economic importance, high-risk supply chains, and the inability of substitution by materials of commensurate properties [14]. According to [15], the approximate amount of antimony exceeds 1.5 × 106 t worldwide. As of 2019, the worldwide reserves of antimony range between 50,000 t, in Tajikistan and 480,000 t, in China, with significant reserves being present in Australia (14,000 t), Bolivia (310,000 t), Mexico (18,000 t), Pakistan (26,000 t), Russia (350,000 t), and Turkey (10,000 t). Antimony is currently virtually non-recyclable [16], due to a number of reasons [17,18], with only a few successful recycling examples existing [19], thus making it a critical material. In the beginning of the 20th century, the major suppliers of Sb were Bolivia and China, later joined by South Africa and the U.S.S.R. [6]. After the 1980s, China showed a rapid and sustained expansion of Sb mining enterprises, and is now responsible for about 87% of the global production [20]. Both heavy metals and metalloids present a particular danger to human health, since they are not biodegradable, and are therefore prone to accumulation in biological systems [21]. A particular issue with heavy metals is that, in contrast to other toxins, they are not destroyed but rather recycled constantly through the geosphere and the biosphere [22– 24]. Antimony toxicology is relatively well documented, owing to its presence in urban environments, its use in a small number of drugs and the relative studies on occupational Int. J. Environ. Res. Public Health 2022, 19, 4669 3 of 30 exposure. In fact, the atmospheric Sb compounds from anthropogenic sources are steadily increasing [25–27], thus predisposing a significant part of the population to potentially associated pathologies. In this review, we will concisely examine the geological processes associated with antimony ore deposits and the basic characteristics of important antimony minerals. Subsequently, we will present the current research on exposure to antimony from an environmental and an anthropogenic standpoint, both at mining sites and in urban settings. We will then examine the toxicology of antimony, the associated pathologies in every major physiological system, and the relevant mechanisms of pathogenesis. Finally, we will go through the use of antimony in medicine, more specifically in the treatment of tropical diseases, and discuss briefly the emerging problem of antimonial resistance. 2. Mineralogy, Geochemistry, and Availability of Antimony It is necessary to understand the geological properties of antimony in order to properly assess the impact of related geological processes on public health [28,29]. Antimony exists naturally in the Earth’s crust and is released into the environment via both natural and mining–industrial processes. Based on current research [30], Sb levels are below 1 mg/kg in rocks and soils, and 0.1 mg/kg in flora and waters. It exists in two biologically relevant oxidation states [31]: the pentavalent form of antimony is common under aerobic conditions, and the trivalent form is common under anaerobic conditions. The availability of antimony in the environment can be regulated by a host of different technological processes, including but not limited to, coagulation, membrane separation, ion exchange, adsorption, and phytoremediation, as presented in [32]. The most common pentavalent-type Sb is found in seawater and freshwater, while antimony released from anthropogenic activities will be most commonly trivalent. The speciation of antimony, from Sb(III) to Sb(V) has already been documented by [33,34]. In the pentavalent state, antimony is very stable and forms complexes with numerous ligands [35,36]; conversely, its trivalent form is a weak base of electropositive character [37]. Antimony itself is a white lustrous metal of average hardness. Geochemically, it is classified as a chalcophile element, meaning that it occurs along with sulfur and Cu, Pb and Ag [13]. There exist over 100 minerals containing antimony [38], belonging to various mineral classes; only a few, however, are important from an economic standpoint (Table 2). Table 2. Economically important antimony minerals. Mineral Stibnite Jamesonite Valentinite Senarmonite Chemical Formula Sb2O3 Pb4FeSb6S14 Sb2O3 Sb2O3 Crystal System Orthorhombic Monoclinic Orthorhombic Cubic (Isometric) Mineral Group Sulfides Sulfosalts Oxides Oxides Stibiconite (Sb3+Sb5+)2O6(OH) Cubic (Isometric) Oxides Bindheimite Pb2Sb2O6O Cubic (Isometric) Oxides Kermesite Tetrahedrite Sb2S2O Cu6(Cu4C22+)Sb4S12S Triclinic Cubic (Isometric) Sulfides Sulfosalts Color Gray with luster Gray to black White to light grey to yellow Colorless to grey White, yellow, orange to light brown Yellow to brown to greenish brown Red Various shades of grey References [39,40] [41,42] [43,44] [45,46] [47,48] [49,50] [51–53] [54,55] The principal antimony ore exists in the form of stibnite (Sb2O3), a sulfide mineral macroscopically appearing as columnar or needle-shaped crystals. The color of the crystals is most commonly a silvery to dark grey, although tarnished crystal faces may have an indigo blue coloration. Stibnite forms in hydrothermal systems and is associated with cinnabar, quartz, and fluorite [55]. Antimony occurs in a number of different deposits such as Sb-bearing minerals (boulangerite, a lead-rich mineral), bournonite [56], gudmundite (an iron-rich mineral), and polybasite, which have also been recognized as minor Int. J. Environ. Res. Public Health 2022, 19, 4669 4 of 30 antimony sources [57]. Antimony ore is frequently associated with undesirable elements, such as Hg [58]. Other accessory minerals are also associated with antimony ores, as mentioned in [57]. According to [57], antimony ores are associated with a number of different deposits, namely epithermal deposits, pegmatite deposits, and hot spring-related replacement deposits. Depending on the ore grade, antimony ore deposits can be categorized as either primary or secondary; in this second category, Sb is mined as a by-product. Sb deposits may be simple (Sb-rich) or complex polymetallic [59]. They can form in many geological settings, and examples of Sb deposits and prospects can be found in the literature [60–73]. A detailed account of the currently active Sb deposits is available in [56]. 3. Exposure to Antimony Exposure to antimony is more frequent in industrial and mining settings but it is also possible during everyday life. At any rate, the recommended maximum exposure level to antimony—also abbreviated as total daily intake (TDI)—is 0.6 μg per kg of body weight per day, as proposed by the WHO [74]. More detailed results, describing different natural and anthropogenic Sb levels, are included in recent research [75]. Extensive use of Sb caused by rapid industrialization and urbanization of the environment rapidly transformed the geochemical character of the soil of many areas [76], mainly due to indirect and direct pollution. A further compounding factor is that the content of such metal pollutants in the soil remains generally elevated, even decades after the removal of the polluting factor, since metals and metalloids exhibit prolonged soil residence times, as described in [77]. In the atmosphere, Sb also presents a potential danger in certain settings; atmospheric Sb compounds are attributable to waste incineration, the use of fossil fuels in internal combustion engines, and road traffic [78,79]. In this part, we will present the contamination and exposure at the different levels, where Sb is introduced into the biosphere due to anthropogenic activities. 3.1. Environmental Contamination and Exposure at Industrial, Mining, and Urban Settings In appreciating the degree of land contamination, it is important to quantify the degree of uncertainty in the methods used to assess soil contamination. A suitable model is that of a probabilistic determination which allows for the determination of an uncertainty factor, as proposed in [80]. As mentioned, industrialization and urbanization triggered a rapid increase in heavy metal content in the biosphere. Soil is regarded as a major reservoir for potentially toxic elements [81]. In general, regarding pollutants, and in particular, antimony, the geochemical character of the soil greatly affects soil chemistry and therefore pollution levels; relevant examples are described in [82,83]. The importance of the geochemical profile of the soil cannot be understated. For example, in a study examining the pollution around the area of Lavrion, Greece, it was found that the availability of toxic elements was limited due to their sequestration in stable mineral phases [84]. This is very important, due to the extensive mining of the carbonate-replacement ore deposits of Lavrion [85], both in antiquity [86,87] and in modern times [88]. If, after aggregate centuries of mining, it is possible for the soil to ‘absorb’ some of the pollutants, then this creates novel opportunities and challenges in assessing industrial heavy metal contamination in general, and Sb contamination in particular. Mining areas are particularly polluted with heavy metals, and only a small portion of heavy metal content can be attributed to geological processes, notably weathering and pedogenesis [84]. Air quality tests indicate that, while some areas with active mining enterprises enjoy reasonable air quality (e.g. [89]), in the majority of cases air quality is negatively affected. In fact, the environmental pollution, which is attributable to mining enterprises, has already been documented by numerous researchers for both ancient (e.g., [90]) and modern mining sites (e.g. [91–95]). Int. J. Environ. Res. Public Health 2022, 19, 4669 5 of 30 Most of the antimony will enter the environment due to mining and industrial activities [96–98]. Indeed, the levels of antimony around smelter sites are exceedingly high, a fact already noted in [99–101]. A particular role is played by waste incineration [102,103] and internal combustion engines. The use of fossil fuels aggravates the situation, given that coal typically contains Sb [103,104]. In contrast to the reference levels provided by [30], in mining areas, soil Sb concentrations can be over 2 × 103 mg/kg and groundwater concentrations over 6 × 103 μg/L. Plant levels may reach up to 143.69 mg/kg [105]. China’s largest deposits are the Xikuangshan Sb ore field, the Dachang Sn-Pb-Zn-Sb ore field, the Zhazaixi Sb deposit, the Xiangxi Au-Sb-W deposit, but of course, many others produce considerable amounts of Sb every year. Extensive research has been performed all these years in most of the mining sites of China (see [106] and references therein), and the results showed significantly increased Sb contents in the soil, water, and plants of most of the mining areas; more specifically, ref. [107] mentions that in the Xikuangshan area, Sb concentrations in the water exceed 53.6 μg/L. Similar examples of Sb contamination are recorded from mining sites in Massif Central, France [108]; Dúbrava, Slovakia [109]; Su Sergiu, Sardinia, Italy [110]; Glendinning, Scotland [111]; Endeavour Inlet, New Zealand [112]; Barcelona [113] and Zamora [114], Spain; Keramos, Chios Island, Greece [115]; and multiple sites in Poland [116–118]. Of course, this is only a small fraction of the recorded cases, and given the current research trend, more papers on the subject will surely come up. Antimony values, in soil and water samples for some of the mining areas referred to above are provided in Table 3. Table 3. Soil and water Sb pollution in selected Sb mining sites. Location Region, Country Sb Water Content (μg/lt) Ouche Pernek Dúbrava Medzibrod Poproč Čučma Su Sergiu Glendinning Endeavour Inlet Llorenç d’Hortons (industrial site) Losacio-Las Cogollas Bardo Bystrzyca Górna Czarnów Dębowina Dziećmorowice Srebrna Góra Puqing mining area Huangshi Xikuangshan Keramos Massif Central, France Malacky, Slovakia Žilina, Slovakia Banská Bystrica, Slovakia Košice, Slovakia Košice, Slovakia Sardinia, Italy Dumfries & Galloway, Scotland New Zealand 200–350 1–31 4–9300 11–1290 5–1000 1–3540 23–1700 0.10–783 14.1–30.4 Sb Soil Content (mg/kg) n/a 121–894 4.8–9619 2–793 13–6786 6.2–782 19–4400 6.77–261 18–243 Barcelona, Spain 1.93–2.06 0.1–112 [113] Zamora, Spain n/a 0.14–0.76 0.13–123 0.01–16.6 0.33–437 0.05–151 0.02–170 n/a n/a n/a 115.94–478.63 60–230 n/a n/a n/a n/a n/a n/a 0.49–1431 0.62–4.65 100–5045 n/a [114] Lower Silesia, Poland Guizhou, China Hubei, China Hunan, China Chios Island, Greece References [108] [109] [110] [111] [112] [118] [106] [119] Regarding the occupational exposure to antimony, a notable problem, concerning its quantification, is that it frequently coexists with other toxic elements, such as As and Pb. As such, it may be difficult or even impossible to separate between the different toxicities [120]. Be that as it may, if protection standards are maintained, save in cases of accidents, the danger is minimal; in countries where, for a host of reasons, safety regulations are lax [121], the danger is increased. Int. J. Environ. Res. Public Health 2022, 19, 4669 6 of 30 Antimony also accumulates in plants, where it enters, at a cellular level, using a number of different aquaporins [122,123]. It is known that there exist numerous Sb species, both stable and unstable, in aqueous environments [25,124,125]. The trivalent form will be predominant in reducing to mildly reducing conditions [126]. The pentavalent form exists in oxidizing environments, such as soils [127–129]. However, Sb(III) prevalence— this is the most toxic species—was recorded in a few cases in [130,131]. In any case, direct human contact with contaminated soil is one of the major exposure pathways, as outlined in [129]. In general, antimony bioavailability in plant-dominated environments exhibits notable variability, e.g. [100,132–138]. Even though antimony is non-essential to plants, it can be taken up through their roots if and when it is available in water-soluble forms [101]. Some plant species, notably Achillea ageratum, Plantago lanceolata, and Silene vulgaris, accumulate antimony readily [101]. For example, according to the results of the aforementioned research, plant concentration can exceed 440 ppm, depending on the plant species and the uptake mechanism. At any rate, Sb accumulation in plants remains pronounced around mining areas [79,139]. In urban settings, a study from Athens, Greece [140], indicated that As is prominently accumulated in parks and woodland areas within the city; taking into account that inorganic As and Sb exhibit similar chemical behavior, it is possible that further research will reveal Sb enrichment in these areas. More prominently, Sb is released into the air by the burning of fire retardants [141], and by brake abrasion particles [142], released during car braking. Recent research has revealed that Sb serum concentrations are elevated in children younger than 6 years in age, in Bucharest [143]. Given the fact that as far as European cities go, there are far more polluted ones, further research is required to establish the susceptibility of young children to heavy metals in urban settings. 3.2. Exposure to Antimony Related to Water Consumption The first obvious source of antimony intoxication would be through the tap water, but the concentration of antimony is usually well below the accepted limit of 1 μg/L [25]; exceptions to this fact are some isolated reports from environmental agencies [144]. A large part of Sb in drinking water is eliminated via water sanitation methods, the most efficient and cost-effective of which is coagulation–flocculation and adsorption [145]. Other methods for achieving the same purpose have also been proposed [146–149]. Despite the efficacy of such methods, the presence of natural organic matter in the water increases the risk of human exposure [150], a fact already reported by [147,151]; the total carbon content is an additional negative modifier for Sb clearance from potable water [152]. The negative influence of natural organic matter can be explained by the formation of Sb–organic matter complexes [153]. Sb binds preferentially with hydrophobic ligands, rather than hydrophilic ones, and the binding potential is higher for Sb(V), compared to Sb(III); the presence of Fe reduces this potential [150]. The dominant species is the pentavalent form of antimony, as reported by [154]. According to the study of [144], the predominance of Sb(V) is explained by the oxidizing agents used during potable water processing, and by the limited stability of Sb(III) in aqueous solutions. Despite reports of no Sb contamination in the overwhelming majority of tap water supply, the situation is rather different for bottled water. In the plastics industry, antimony trioxide (Sb2O3) is used as a catalyst in polyethylene terephthalate production (PET); consequently, the plastic making up the bottles contains between 190–300 mg·kg−1. Due to antimony leaching, the amount of antimony in the water is directly proportional to the duration of plastic water bottle storage [155]. This problem is further aggravated by the increase in bottled water consumption during the last decade [156–160]. The storage of plastic water bottles at higher temperatures increases antimony leaching, further contaminating the water contained within [155,161–163]. Int. J. Environ. Res. Public Health 2022, 19, 4669 7 of 30 3.3. Exposure to Antimony Related to Food Consumption The entry point of Sb into the food chain is through plants, which absorb it from contaminated soil. However, the degree of antimony soil contamination is not the sole determining factor, as its mobilization in the soil greatly affects its uptake by the local flora; this has been demonstrated in [134,139]. Another modifying factor is the position of the plants in the food chain, i.e., if the plant is consumed directly by humans, or if it enters the food chain after being consumed by herbivores. This was demonstrated in [164–166] for mushrooms and radishes. Some data on the presence of Sb in milk exist [167–170]. The intake of antimony from milk was calculated to be less than the limit applicable in the case of water, but the comparison between the results of different studies is sometimes difficult, due to different calculation methods. Studies on wine samples indicate that most often antimony is below the detection limits of the applied methods; the use of other methods indicated Sb levels close to 10 μg·L−1 in some European wines [170]. Seafood is generally considered not to be a source of contamination in most areas, although data from industrial coastal zones indicate that this is not always the case [171]. It is believed that industrial activity is the direct cause for the results of this study. The predominant species of antimony in seafood is the pentavalent form. But even if food itself is not considered a high-risk source for Sb, its packaging is not exempt as a source for concern. The plastics used in food packaging are manufactured by the same process used for water bottle manufacturing and are therefore prone to contaminating packaged food with antimony, an occurrence which becomes especially pronounced if the plastic container is heated in a microwave oven [144]. Research by [172] also indicates that the presence of citrus juice may increase antimony leakage from the plastic packaging; it is hypothesized that the citric acid preserves the oxidation state of the leached Sb(III), a process already demonstrated in [173]. Even in the latter case, however, the antimony content of beverages was below the acceptable levels. 4. Toxicity and Toxicology of Antimony In general, heavy metal toxicity affects negatively the body’s systems, and long-term exposure will lead to the appearance of degenerative phenomena. The toxicity of antimony is monitored by assessing various environmental factors [174]. In addition, not all Sb which enters the body will participate in adverse reactions. Rather, only a fraction of it, attached to water molecules or various particles, will enter through the respiratory and/or the gastrointestinal tract [144]. The current trend in assessing human exposure to pollutants has shifted towards calculating the bioaccessibility of each pollutant and not just its total content [175–177]. An example of such an application was illustrated for Pb by [178]. The toxicity of antimony is regarded to be on par with, or even higher than, that of arsenic, and the inorganic form of antimony, Sb(III), is far more toxic, than the organic one, Sb(V). In general the inorganic species are most often the more potent forms in terms of toxicity. Despite antimony’s similarities to arsenic, it should be noted that only the biochemical behaviors of the trivalent form are comparable [168]; the pentavalent forms of antimony and arsenic have different structures [179] and may thus affect different physiological mechanisms of the human body [4]. According to [180], the toxicity of antimony is derived by its binding to thiol-containing enzymes. Specifically, the organic form of antimony is almost harmless to red blood cells, since it cannot penetrate their cell membrane, while the inorganic form shows a high affinity both for red blood cells and thiol groups. As antimony complexes with thiol groups, forming thioantimonites, it is presumed that the GSH levels within the cells are depleted, an effect already observed during exposure of cells to As [181], which exhibits comparable chemistry and toxicity, and also complexes with thiol groups. It is also not Int. J. Environ. Res. Public Health 2022, 19, 4669 8 of 30 improbable that the thiol groups of some proteins interact with Sb in a similar manner to the thiol groups of glutathione. Glutathione peroxidase is also affected negatively by Sb, and this decreases free GSH levels, leaving the cells yet more susceptible to oxidative stress [182]. Sb and As are direct inhibitors of pyruvate dehydrogenase, the basic regulatory enzyme determining the mode of glucose oxidation, i.e., anaerobic or aerobic. Exposure of cells to antimony leads to an observed drop in ATP levels, and it is hypothesized that the inhibition of pyruvate dehydrogenase by Sb, activates the anaerobic glycolysis pathway. Anaerobic glycolysis is, of course, vastly more inefficient than aerobic glycolysis and produces far less ATP. It is also hypothesized that the trivalent form of antimony is potentially carcinogenic [4]; currently, it is regarded as being carcinogenic for animals [183]. Up until recently, the mutagenic and carcinogenic potential of antimony had received meager attention, compared to studies on other heavy metals such as Pb, e.g., [184–187]. A recent study [187] ascertained that there is a probable correlation between the level of antimony trioxide inhalation in Sb smelter workers and detected DNA lesions. The mechanisms associated with Sb-mediated DNA damage are discussed in [188,189] and presented in a concise way in [75]. It is interesting to note that DNA damage was analogous to the urinary antimony levels; this study concurs with the findings of the only other significant study of Sb genotoxicity, [190]. Some researchers have performed animal trials, to determine if Sb is genotoxic, but the results were either negative [191] or marginally positive; thus, there is as of yet no consensus in the scientific community. Nonetheless, in vitro experiments in cells proved that Sb can cause cell death [192], inhibit DNA repair mechanisms [193], and interfere with transcription mechanisms [194]. Currently, the prevailing hypothesis is that Sb genotoxicity is mostly caused by its interference with repair mechanisms [195,196]. Antimony also interferes with the metabolism of sugars in the human body and can bind to many of them [197]. Its trivalent form inhibits gluconeogenesis [75] and promotes the pentose phosphate pathway [198]. Imbalances in lipid metabolism, caused by Sb, may also enhance its carcinogenic potential [199,200]. Sb(III) has been also linked to a hemolytic mechanism by [201]. Regarding the effects of Sb on the reproductive capacity of humans, there are as of yet no definitive conclusions. At first glance, the most severe effects seem to be associated with the increased mutation rates caused by Sb-related DNA damage leading to abnormal genotypes in offspring. According to [202], Sb exposure is linked to decreased sperm count, although based on relevant research [203], there is no decrease in semen quality, at least when Sb concentrations in the plasma are low. Moreover, in pregnant women, Sb accumulation is linked to increased incidence of pregnancy-induced diabetes mellitus [204,205] and perhaps hypertension [206]. Other pregnancy-related and development-related risks are described in [75]. 4.1. Cellular Mechanisms Associated with Antimony Entry and Processing Antimony enters the cells via aquaporin channels [207–209]. More specifically, it has been proven that Sb(III) enters the cells through the GlpF aquaporin channel in Escherichia coli, the same channel that mediates the entry of As(III) into that organism [210,211]. The GlpF protein belongs to the sub-family of aquaglyceroporins, because it allows not only water, but small uncharged solutes to pass through [212,213]. Later [213] proved that the Fps1 protein of the same sub-family is responsible for the entry of Sb(III) in Saccharomyces cerevisiae, thus illustrating the entry pathway in a eukaryotic cell for the first time. Because these proteins exist in cells of all gena and species, it can be said with a measure of certainty that this is the evolutionarily conserved pathway for the entry of metalloids into cells [207–209]. For humans, AQP9 is implicated in antimony transport, according to recent experiments [214,215]. Given the bidirectionality of aquaporins [122,216,217], it is probable that they may transport Sb out of the cell too, thus acting also as a detoxification mechanism. Based on the fact that the GLUT1 transporter [218,219] and hexose permeases Int. J. Environ. Res. Public Health 2022, 19, 4669 9 of 30 can catalyze As(III) transport [197] in some non-human cells, it can be hypothesized that such proteins might be implicated on Sb(III) transport as well. The route of entry of Sb(V) remains unknown [4]; perhaps a clue may lie in the phosphate transporters which have been shown to transport As(V) [208,220,221], but this is a contested issue given the differences in the biochemical character of this particular oxidation state of these two elements. Regarding Sb reduction intracellularly, the only Sb-specific mechanisms are known from unicellular organisms of the Leishmania species [222–225]; potentially some correlation to human cells can be made in the future. A more general mechanism, which was initially hypothesized as one of the causes of toxicity by [182] is the interaction between Sb and glutathione, as analyzed in [222]. 4.2. Physiological Mechanisms of Sb Toxicity Reduction in the Human Body As mentioned above, antimony is potentially carcinogenic, but some of its toxic potential is reduced by a host of cellular mechanisms, which decrease its cytosolic content. Cells can limit Sb import, force its export, sequester it in intracellular organelles, or possibly chelate it [226–229]. It has been proposed that the rapid expulsion of antimony from the cells might be related to the development of resistance to it. The reduction of antimony from its pentavalent to its trivalent form is the principal mechanism behind the action of antimonials against leishmaniasis [230]; this will inhibit the action of glutathione and trypanothione. It can potentially then be expelled via the As pump [231]. Recent research [222] indicates that the reduction of antimony toxicity is regulated by the availability of glutathione. Glutathione catalyzes, via a redox reaction, the chemical reduction of Sb(V) to Sb(III), in a dose-dependent manner. This reaction happens faster in acidic pH values and at higher temperatures. Based on a similar redox reaction between glutathione and As [232], it is reasonable to assume the creation of an SbGS3 complex, a probability further supported by the findings of [233]. The total oxidation reaction may be written thus: SbO + H + 2GSH ↔ HSbO + GS-SG + H O (1) where GS-SG represents the oxidized form of glutathione. The thermodynamical parameters of this reaction are presented in [234]. Antimony is expelled by the human body through renal filtration and excretion in urine [235], with different excretion rates recorded in China and Sweden by [236] and [237], respectively. Methylation, both by human cells and gut microbiota, has also been linked to Sb neutralization and removal from the human body [238,239]. 4.3. Effects on the Respiratory System There is an incomplete set of data, regarding the absorption of antimony in the respiratory tract. It has been established as a quantifiable occurrence, both from the studies on occupational exposure to antimony and relevant animal experiments. According to [240], the average absorption is 15%, a percentage similar to what occurs in the gastrointestinal tract. Particle size and solubility were considered as the main modifying factors. Several researchers [241–246] have recorded elevated Sb levels in the blood and urine of workers exposed to antimony in mining and industrial settings. Given that the only form of exposure in these studies involved inhalation, it is evident that at least some degree of absorption must happen in the respiratory tract. Another research by [247], specifically on pregnant women working in Sb smelters, revealed that Sb was detectable in the placenta and the amniotic fluids. A confounding issue was that the levels of Sb in body tissues and fluids were not enough evidence to quantify absorption, and the chance that a portion of the inhaled antimony had been ingested and thus removed from the respiratory tract altogether before absorption, could not be excluded [144]. The systematic research of [248–251] proved that concentrations of Sb were elevated in the lungs of occupationally exposed people, thus corroborating that inhalation is Int. J. Environ. Res. Public Health 2022, 19, 4669 10 of 30 proportional to some degree to the Sb air particle content. Studies on animals have also been performed [241,252–258]. But these suffer from more or less the same constraints presented below for animal examples on antimony absorption in the gastrointestinal tract. This question had already been raised in [259]. Finally, painless ulceration and perforation of the nasal septum was described in [260,261] in occupationally exposed workers. Even though Sb may be the culprit, it is hypothesized that the coexistence of As in these settings is the most probable cause [262]. 4.4. Effects on the Cardiovascular System It has long been recognized that the exposure of mammals to Sb-containing compounds is particularly dangerous for the cells of the myocardium. The first indication of this phenomenon was [263], in experiments with rats. Further research [241] indicated that the administration of potassium antimonyl tartrate increased the degeneration of the fibrous and connective tissue of the heart, even at low doses. The adverse effects of exposure to antimony were also studied in the exams and autopsies performed on patients who had received antimonial drugs; Sb was identified as being the cause of death, due to its specific toxicity to the heart, an effect observable even in the altered form of some electrocardiograms [264–266]. The earlier study of [241] had correlated Sb-induced heart problems, also detectable by abnormal electrocardiograms, with the death of workers exposed to antimony trisulfide for a relatively prolonged average period. It has been already proven that antimony increases the oxidative stress in myocardial cells [267], and the subsequent experiments of [182] proved that in vitro, myocardial cells exposed to Sb exhibited increased cell death incidence. This is tied to the presumed decrease in GSH levels and the interdiction of Sb in the activity of certain enzymes, as described above. The death of myocardial cells must also be associated with the observed drop in ATP levels, in cells exposed to Sb [182]. 4.5. Effects on the Oral Cavity Antimony is among a variety of metals that can be detected in the oral cavity and may be introduced as part of various dental materials [268]. A recent study using photoactivation analysis found trace elements of antimony alongside nickel, barium, arsenic, strontium, and others in dental composites manufactured by various producers [269]. The long-term release of antimony from dental materials might cause chronic exposure with severe effects. A study [270] using cell viability assays has shown that antimony demonstrates weak embryotoxicity, a finding that correlates with previous similar reports [271,272]. Furthermore, the presence of antimony in the oral cavity can induce a change in the salivary microbiome composition. In their recent study, the authors of [273] proposed that the presence of salivary metals will induce changes in the oral microbiome and lead to oral health issues. In subjects with increased antimony levels, they reported a higher abundance of Lactobacillus and Granulicatella species, which have been associated with the development of dental caries and increased dental decay [274,275]. Conversely, a recent study concluded that electronic cigarette smoking is not a source of increased antimony levels in the body, as the tested urine of regular e-cigarettes smokers showed similar levels of antimony and other heavy metals to that of persons who never used electronic cigarettes [276]. 4.6. Effects on the Gastrointestinal Tract The absorption of antimony through the gastrointestinal tract is estimated to be about 5–15% of the total amount of Sb ingested [240,277,278]. It has been established that antimony causes notable side effects when introduced to the gastrointestinal tract. Its use as an emetic has already been mentioned. There are some cases of oral poisoning reported [246,279,280] which indicate, firstly, that up to a point, Sb Int. J. Environ. Res. Public Health 2022, 19, 4669 11 of 30 is absorbed in the gastrointestinal tract, and secondly, that it is poisonous even if we accept that the maximum absorption is only 15% of the ingested value. A single case study [246] reported that antimony was not detectable in the gastric juice and bile after 100 h, but serum and urine levels remained abnormally high even after one week. There have been numerous in vivo studies in animals [255,281–285] that have yielded a maximum of 18% gastrointestinal absorption of antimony tartrate. It was observed, however, that there were significant differences regarding absorption based on the delivery method. Intraperitoneal injection of the drug proved lethal whereas the oral administration was not, due to the poor gastrointestinal absorption of antimony [285]. Lastly, the authors of [286] conducted experiments regarding the Sb levels in the red blood cells of rats and observed that the Sb concentration was dose-dependent and higher in female rats. Further experiments [254,287–292], while presenting somewhat different results, corroborate that there is at least a degree of absorption. Despite ample evidence from animal trials, the results cannot be easily correlated to humans due to some significant constraints, outlined in [144]: the administered forms contain much more Sb than would happen in realistic conditions; the chemical forms of the ingested Sb are not frequently found in nature; and in all organisms, other dietary and health factors affect absorption and tolerance of antimony. 4.7. Effects on the Skin What few data exist on Sb and skin interaction originate from studies on smelter workers and miners. The first such research was performed by the authors of [293], who recorded skin irritations. The so-called ‘antimony spots’, i.e., antimony-associated skin lesions, were recorded in [294]. Antimony exposure has also been linked to occupational dermatitis by some researchers [295,296]. In any case, the appearance of antimony spots is considered a rare occurrence [262]. From a histological perspective, these lesions exhibit necrosis and acute inflammation closely related to the sweat ducts [294]. The lesions, which often exhibit eczema and lichenification, resemble those of smallpox; they occur in many body locations, except for the feet, hands, and face [262]. An interesting curative use of antimony was its use in Mohs paste [297], which was developed in the 1930s by F.E. Mohs [298], who observed that a 20% solution of zinc chloride caused cellular death but preserved the general histological structure [299]. The application of this paste, albeit with an altered composition, is the basic step of the still-inuse, fresh-tissue Mohs chemosurgery technique, proposed in the 1960s, to rectify the increased incidence of recurrence of basal cell carcinomas after the initial surgery had taken place [300]. The Mohs chemosurgery was found to be highly effective in the case of basal cell carcinoma [300] as well as other rarer cases, such as the excision of cylindromas [301]. The advantages of zinc chloride, the basic component of Mohs cream, are that it is a good fixative and its permeation can be controlled through the application of a paste of specific composition [302]. The stibnite served as the granular part and sanguinaria canadensis, commonly known as bloodroot powder, served as the powder. It is worth noting that Mohs tested a number of compounds as in situ fixatives, including antimony trichoride, which, however, distorted tissue structures [303]. Even so, the application of the paste was painful and sometimes caused lymphadenopathy; local inflammation and fever were also not uncommon [303]. Such effects are more associated with the toxicity of bloodroot, whose other preparations have similar or even more adverse side-effects [304]. Such problems were obviated with the introduction of the fresh-tissue technique, where no paste is applied. The current iteration of the paste, which has a number of applications, as discussed below, uses neither stibnite nor bloodroot. The clinical application of the cream began in 1936, following successful in vivo trials in rats; the initial applications of the method were highly successful [305–308]. Gradually the fresh-tissue Mohs chemotherapy was developed which led to ever more efficient tumor excisions (e.g., [309–312]). It must be noted, however, that the high success rates of Int. J. Environ. Res. Public Health 2022, 19, 4669 12 of 30 the technique are attributable to the fixative and not the anti-cancer properties of the paste [313], as demonstrated by [314]. Of course, the process also has some drawbacks, in that it was painful and time-consuming and the devitalization of the tissues made the closure of the incision difficult [315]. Today, fresh-tissue micrographic surgery, which does not use the paste, is much more rapid and less discomforting to the patients. Apart from chemotherapy, the Mohs paste has proven useful in a number of occasions and clinical settings (e.g., [315–319]), although stibnite is no longer used to provide the granular part; rather the current composition of the paste comprises zinc chloride, distilled water, zinc starch, and glycerol [320]. 5. Use of Antimony in Medicine While, as mentioned before, the uses of antimony can be traced back to ancient civilizations, it was Paracelsus who promoted its use during the 17th century in Europe. However, the systematic use of antimony in modern medicine can be attributed to Plimmer and Thomson, who used it to treat African trypanosomiasis [321]. Soon, the side effects of antimonial drugs, which include but are not limited to headache, nausea, vomiting, diarrhea, ache in the muscles and joints, coughing and syncopes, and anaphylaxis, became apparent [322]. Interestingly, there is a single, apparently positive report, in using antimonials to treat syphilis [323]. Some attempts were also made to use Sb against malaria [324,325] and at least one attempt was made to treat framboesia tropica (non-venereal endemic syphilis) [326]. Currently, the use of antimony in the treatment of lung tumor cell lines [120] is being studied. In addition, quite recently, the use of Sb dithiocarbamate complexes has been studied for their potential antibacterial activity, with promising results [327], and they have also exhibited a noteworthy antifungal activity [328]. Based on the general anticarcinogenic principle of action of dithiocarbamate compounds, Sb-dithiocarbamate compounds can be considered as potential anticarcinogenic agents [329]. The anticarcinogenic potential of Sb has already been mentioned in relation to the research of [120]. A summary of the most relevant applications of antimony in medicine is presented in Table 4. Table 4. Summary of medical uses for antimony compounds. Pathology Cancer Syphilis Malaria Compound and Administration Trivalent antimony potassium tartrate Antimony powder in saline solution—intravenous injections Various Dosage 4.2–322 μg/mL Pathogenic Factors Application Targeted small cell lung canin vitro cer cell lines (currently under research) 50–200 mg Treponema pallidum Various Plasmodium spp. Framboesia trop- Antimonium tartarum— ica intramuscular Various Sb(ephedtc)3 and monoVarious bacterial phenylantimony(III) infections compounds—microtiter plates & salt application 21.4–125.6 μM in vivo (historical use) in vivo (historical use) in vivo (historical use) Treponema pallidum pertenue P. aeruginosa; E. coli; K. pneumoniae; Salmonella dublin; E. cloacae; S. aureus; in vitro E. caseofluvialis; S. (currently under research) sciuri; plus multiresistant clinic isolated strains References [330] [323] [324,325,331] [326,332] [327,328] Int. J. Environ. Res. Public Health 2022, 19, 4669 13 of 30 Monophenylantiin vitro mony(III) compounds— 27.9–65.08 μM A. niger; A. flavus [328] (currently under research) Salt application Sodium antimony gluLeishmaniasis conate; meglumine anti10–100 mg/kg Leishmania spp. in vivo [321,333–335] moniate—intramuscular Various combinations of in vitro (experiments in muTrypanosomiasis antimonials and other Various Trypanosoma spp. [336–339] rine trypanosomiasis); in vivo compounds Various antimonials— in vivo Schistosomiasis intravenously, intramus3.5–530 mg Schistosoma spp. [340–349] (historical use) cular ephedtc = ephedrinedithiocarbamate ligand. Aspergillosis 5.1. Antimonial Drugs for Leishmaniasis Treatment Leishmaniasis can be regarded as a complex zoonosis and is caused by protozoans of the genus Leishmania. About 20 species of Leishmania are infectious to humans, and their vectors are different species of female phlebotome sandflies [350,351]. Currently, it is considered endemic in Africa and Asia, but even in Western countries, it is a problem in HIVinfected patients [352], or patients who are otherwise immunosuppressed for medical reasons. The partial or even complete lack of a functional immune system leads to increased parasite burdens and compromised treatment response [353]; such patients are more liable to develop antimonial drug resistance [354,355]. According to [352], patients with concomitant AIDS and leishmaniasis can infect sandflies which will further spread the disease; the same is not true for immunocompetent patients. This emerging problem can be partially mitigated by the use of anti-retroviral drugs, a course of treatment that is, however, not universally available [356]. The most common use of antimonial drugs in medicine concerns in the treatment of leishmaniasis. There exist two main types: trivalent antimonials, also known as tartar emetics, and pentavalent antimonials. The first confirmation of antimonial drug efficacy was provided in [357] against cutaneous leishmaniasis, and in [358,359] against visceral leishmaniasis. Despite the initial hopeful results, the toxicity of the drugs soon became evident; a problem additionally compounded by their apparent instability in tropical climates [360], where the disease is most prevalent. Other reports, however, refs. [361,362] indicated that the use of tartar emetics was ineffective to the point that it was proposed that, perhaps, no treatment would be preferable. A symptom of cutaneous leishmaniasis are the so called ‘oriental sores’ and reports of Sb use in their treatment is mentioned in [363,364]. Nowadays, the classic therapy for leishmaniasis is pentavalent antimony, to which there appears to be, however, increasing resistance. In cases of such an occurrence, liposomal amphotericin B is preferred, which is, however, much more expensive. It has been observed, by various researchers over the years [365–370] that trivalent Sb compounds are toxic to both stages of Leishmania parasites, i.e., the amastigotes occurring with the mammals and the promastigotes occurring in the sandflies; by comparison, pentavalent antimonials are only toxic to amastigotes. 5.2. Antimonial Drugs for Human African Trypanosomiasis Treatment Human African trypanosomiasis is a serious condition, which will prove fatal if left untreated. The number of worldwide reported cases is remarkably low, but its regional distribution and localization are pronounced in sub-Saharan Africa. The pathogenic species responsible for the disease are the unicellular protozoans Trypanosoma brucei gambiense, and Trypanosoma brucei rhodesiense; while Trypanosoma brucei brucei infects only animals. Tsetse flies of the Glossina genus are responsible for the transmission of Int. J. Environ. Res. Public Health 2022, 19, 4669 14 of 30 Trypanosomes [371]. The pathogenic Trypanosoma species are immune to the actions of the human physiological trypanosomal lytic factor [372]. The choice of drugs for treating human trypanosomiasis is rather limited. For the first stage of the disease caused by the T.b. gambiense, pentamidine is used, either intramuscularly or intravenously [373,374]. For the intermediate stage of the disease, pentamidine has proven rather ineffective. Being a tropical disease, antimony was once a considered a prime candidate as a potentially curative agent [336]. Today the use of antimonials is sparse, owing to their very high toxicity, both in cases of humans, and when used to treat animal trypanosomiasis [375]; despite that, in vitro results [338] and in vivo studies [337,339] of some antimonials have proved positive. 5.3. Antimonial Drugs for Schistosomiasis Treatment Schistosomiasis, also known as bilharziasis, is the second most prevalent tropical disease after malaria [376], and is common in tropical and subtropical regions [377–379]. Schistosomiasis is caused by worms of the genus Schistosoma [379], with water snails being the intermediate hosts of the parasite. Schistosomiasis can be distinguished between acute, also known as Katayama fever, and chronic [376]. Drugs based on antimony inhibit glycolysis and other metabolic pathways [362]. In 1918, sodium antimony tartrate began being used for the treatment of schistosomiasis [340,341] and it was found to be quite effective until the 1960s [380–382] under different treatment protocols. Different antimony drugs, or in different doses, were also used in [342,343]. An attempt in [344] to use an oral antimony salt had disappointing results and it was concluded that trivalent sodium antimony tri-gluconate was ineffective when administered orally. Comparative trials of different antimonials demonstrated, in the case of urinary schistosomiasis at least, that the most effective drugs were accompanied by the most severe side effects [345]. A new chemical form of the standard antimony sodium tartrate was proposed in [346] in the early 1970s, and according to an experiment conducted by the researchers, it was characterized by better tolerance. A few years before, in 1968, a very successful use of sodium antimony tartrate was reported in [347], for patients specifically infected with S. haematobium. Today, antimony drugs are no longer used to treat schistosomiasis, mainly because of their cumulative toxicity and the fact that the maximally active Sb(III) linked to oxygen species was very toxic, while the less-toxic Sb(III) linked to sulfur species was also less active. Other antimonial drug formulations were attempted [348,349], but towards the 1970s, the use of Sb in treating schistosomiasis was abandoned. From an early stage, special consideration was given to the adverse effects of antimonials to the myocardium. Initially, there were some reports of death shortly after [383] or sometime later [384], following the administration of trivalent antimony compounds. Further researchers also noted anomalies in cardiograms of treated patients and heart-related pathologies ([264] and references therein). Based on these reports, the authors of [264] conducted research on the electrocardiograms of the treated patients, reporting slight changes in most patients, while in a number of patients the changes were so severe as to indicate myocardial disease induced by the treatment. While the degree of the changes could not be correlated with the dose, in each individual they became progressively worse during the course of the treatment. The most prominent changes were in the T wave, and in all cases they diminished fairly rapidly after the cessation of the treatment [264]. 5.4. Resistance to Antimonial Drugs The initial use of antimonial drugs for over half a century did not indicate any notable development of resistance [321], although, perhaps, such observations are difficult to make, given the regional variation of treatment protocols. For example, in most areas of South America, Africa, and Asia, where the disease is prevalent, the standard treatment protocol is a dose of 20 mg per kg each day of a pentavalent antimonial, which is Int. J. Environ. Res. Public Health 2022, 19, 4669 15 of 30 administered parenterally, for a period of about a month. In the Mediterranean region, the treatment of choice is liposomal amphotericin B (L-AmB). Both of the treatments mentioned involve immunocompetent patients [108]. The cardinal rule behind drug resistance is that a drug with the smallest ratio of halflife to therapeutic efficacy has the lowest possible chance of inducing resistance. This explains, for example, the very high efficacy of amphotericin B deoxycholate in the case of leishmaniasis [385,386]. The resistance to antimonials is most probably associated with the detoxification mechanism of the cells, as mentioned above. In the particular case of leishmaniasis, but in other tropical diseases too, the choice of drug depends on efficacy, toxicity, cost, and availability, in that order of importance [387]. It is not currently known if the inducement of resistance to antimonials can be attributed to their extensive use or to their specific properties in terms of their absorption and action inside the human body. 6. Discussion and Conclusions As is apparent from the section on antimony in the food chain, most foodstuffs are considered safe, although the relevant studies are rather limited, and frequently the detection methods employed are not sensitive enough to detect Sb, even though its presence is speculated. Hence, the confidence level of analytical studies remains relatively low [144]. According to the literature review in [388] there is a potential for micronutrients to modulate the adverse effects of heavy metal intoxication, and hence antimony intoxication. More specifically, dietary sufficiency or insufficiency can greatly modulate the risk assessment of such metals. Taking into account the rising public concern regarding the accumulation of prospective harmful elements and other contaminants, both in the geosphere and biosphere [389], detailed investigations into Sb pollution and contamination are required. Some case studies for some pollutant loads exist for rivers (e.g., [390,391]) and other ecological sites. In the specific context of soil contamination assessment, the geological background should always be taken into account [392]; a holistic ecosystem approach, as proposed by [393] might be optimal for this purpose. Further research is required in the future to ascertain the dispersion of Sb, via the atmosphere in urban settings. This is further illustrated by the link between airborne particles and disease in such settings (e.g., [394,395]). Research along the lines of the methodology of [396] would be useful in such an endeavor. Author Contributions: Conceptualization, A.P., A.-T.P. and C.S.; formal analysis, investigation, data curation, and writing—original draft preparation, A.P., A.C., A.-T.P., A.-E.S., I.A.B., C.C. and C.S.; writing—review and editing, A.P., A.C., A.-T.P., C.C. and C.S.; supervision and project administration, I.A.B., C.C. and C.S. All authors have read and agreed to the published version of the manuscript. Funding: This research received no external funding. Institutional Review Board Statement: Not applicable. Informed Consent Statement: Not applicable. Data Availability Statement: Not applicable. Conflicts of Interest: The authors declare no conflict of interest. References 1. 2. Jaishankar, M.; Tseten, T.; Anbalagan, N.; Mathew, B.B.; Beeregowda, K.N. Toxicity, mechanism and health effects of some heavy metals. Interdiscip. Toxicol. 2014, 7, 60–72. https://doi.org/10.2478/intox-2014-0009. (In English) Jarup, L. Hazards of heavy metal contamination. Br. Med. Bull. 2003, 68, 167–182. https://doi.org/10.1093/bmb/ldg032. Int. J. Environ. Res. Public Health 2022, 19, 4669 3. 4. 5. 6. 7. 8. 9. 10. 11. 12. 13. 14. 15. 16. 17. 18. 19. 20. 21. 22. 23. 24. 25. 26. 27. 28. 29. 30. 31. 16 of 30 Morais, S.; Costa, F.G.; Pereira, M.D.L. Heavy Metals and Human Health. In Environmental Health–Emerging Issues and Practice; Oosthuizen, J., Ed.; IntechOpen: Joondalup, Australia, 2012; pp. 227–246. Tamás, M.J. Cellular and molecular mechanisms of antimony transport, toxicity and resistance. Environ. Chem. 2016, 13, 955– 962. https://doi.org/10.1071/en16075. Grund, S.C.; Hanusch, K.; Breunig, H.J.; Wolf, H.U. Antimony and Antimony Compounds. In Ullmann’s Encyclopedia of Industrial Chemistry; Wiley-VCH: Weinheim, Germany, 2006; pp. 11–42. Butterman, W.C.; Carlin, J.F., Jr. Mineral Commodity Profiles: Antimony. In Open-File Report; Report 2003-19; USGS: Reston, VA, USA, 2004. Available online: http://pubs.er.usgs.gov/publication/ofr0319 (accessed on 13 January 2022). Li, T. Antimony and Antimony Alloys. In Kirk-Othmer Encyclopedia of Chemical Technology; John Wiley & Sons: New York, NY, USA, 2011; pp. 1–15. Schrader, F.C. Antimony in 1919. In Metals: U.S. Geological Survey Mineral Resources of the United States; USGS: Washington, DC, USA, 1922; Volume 1, pp. 286–311. Johnstone, P.; McLeish, C. World wars and the age of oil: Exploring directionality in deep energy transitions. Energy Res. Soc. Sci. 2020, 69, 101732. https://doi.org/10.1016/j.erss.2020.101732. (In English) Slot, B.J. The «Original Naxian Emery» in the International Economy (14th–19th cen.). Flea 2008, 19, 17–19. Periferakis, A. The Importance of Emery in the Cultural, Social and Economic Development of Naxos Island, Cyclades, Greece. Presented at the 15th International Congress of the Geological Society of Greece, Athens, Greece, 22–24 May 2019. Periferakis, A. The Emery of Naxos: A Multidisciplinary Study of the Effects of Mining at a Local and National Context. J. NXA Multidiscip. Peer Rev. J. 2021, 7, 93–115. Anderson, C.G. The metallurgy of antimony. Geochemistry 2012, 72, 3–8, https://doi.org/10.1016/j.chemer.2012.04.001. Tzamos, E.; Gamaletsos, P.N.; Grieco, G.; Bussolesi, M.; Xenidis, A.; Zouboulis, A.; Dimitriadis, D.; Pontikes, Y.; Godelitsas, A. New Insights into the Mineralogy and Geochemistry of Sb Ores from Greece. Minerals 2020, 10, 236. Available online: https://www.mdpi.com/2075-163X/10/3/236 (accessed on 13 January 2022). Klocho, K. Antimony. In Mineral Commodity Summaries; U.S. Geological Survey: Reston, VA, USA, 2019. Karlsson, T.; Forsgren, C.; Steenari, B.-M. Recovery of Antimony: A Laboratory Study on the Thermal Decomposition and Carbothermal Reduction of Sb(III), Bi(III), Zn(II) Oxides, and Antimony Compounds from Metal Oxide Varistors. J. Sustain. Met. 2018, 4, 194–204. https://doi.org/10.1007/s40831-017-0156-y. Graedel, T.E.; Reck, B.K. Recycling in Context. In Handbook of Recycling; Worrell, E., Reuter, M.A., Eds.; Elsevier: Boston, MA, USA, 2014; pp. 17–26. Rombach, E.; Friedrich, B. Recycling of Rare Metals. In Handbook of Recycling; Worrell, E., Reuter, M.A., Eds., Elsevier: Boston, MA, USA, 2014; pp. 125–150. Yellishetty, M.; Huston, D.; Graedel, T.; Werner, T.; Reck, B.K.; Mudd, G. Quantifying the potential for recoverable resources of gallium, germanium and antimony as companion metals in Australia. Ore Geol. Rev. 2017, 82, 148–159. https://doi.org/10.1016/j.oregeorev.2016.11.020. European Commission, Directorate-General for Internal Market, Industry, Entrepreneurship and SMEs. Study on the Review of the list of Critical Raw Materials. Criticality Assessments; Publications Office of the European Union: Brussels, Belgium, 2017. Marsan, F.A.; Biasioli, M. Trace Elements in Soils of Urban Areas. Water Air Soil Pollut. 2010, 213, 121–143. https://doi.org/10.1007/s11270-010-0372-6. Sullivan, M.J.; Leavey, S. Heavy metals in bottled natural spring water. J. Environ. Health 2011, 73, 8–13. (In English) Liu, B.; Wu, F.; Li, X.; Fu, Z.; Deng, Q.; Mo, C.; Zhu, J.; Zhu, Y.; Liao, H. Arsenic, antimony and bismuth in human hair from potentially exposed individuals in the vicinity of antimony mines in Southwest China. Microchem. J. 2011, 97, 20–24. https://doi.org/10.1016/j.microc.2010.07.008. Xi, J.; He, M.; Wang, P. Adsorption of Antimony on Sediments from Typical Water Systems in China: A Comparison of Sb(III) and Sb(V) Pattern. Soil Sediment Contam. Int. J. 2014, 23, 37–48. https://doi.org/10.1080/15320383.2013.774319. Filella, M.; Belzile, N.; Chen, Y.-W. Antimony in the environment: A review focused on natural waters: I. Occurrence. EarthScience Rev. 2002, 57, 125–176. https://doi.org/10.1016/s0012-8252(01)00070-8. Krachler, M.; Zheng, J.; Koerner, R.; Zdanowicz, C.; Fisher, D.; Shotyk, W. Increasing atmospheric antimony contamination in the northern hemisphere: Snow and ice evidence from Devon Island, Arctic Canada. J. Environ. Monit. 2005, 7, 1169–1176. https://doi.org/10.1039/b509373b. Iijima, A.; Sato, K.; Yano, K.; Kato, M.; Kozawa, K.; Furuta, N. Emission Factor for Antimony in Brake Abrasion Dusts as One of the Major Atmospheric Antimony Sources. Environ. Sci. Technol. 2008, 42, 2937–2942. https://doi.org/10.1021/es702137g. (In English) Cook, A. Public Health and Geological Processes: An Overview of a Fundamental Relationship. In Essentials of Medical Geology (Revised Edition); Selinus, O., Ed.; Springer: Dordrecht, The Netherlands, 2013; pp. 15–32. Davies, B.E.; Bowman, C.; Davies, T.C.; Selinus, O. Medical Geology: Perspectives and Prospects. In Essentials of Medical Geology (Revised Edition); Selinus, O., Ed.; Springer: Dordrecht, The Netherlands, 2013; pp. 1–13. Reimann, C.; Matschullat, J.; Birke, M.; Salminen, R. Antimony in the environment: Lessons from geochemical mapping. Appl. Geochem. 2010, 25, 175–198. https://doi.org/10.1016/j.apgeochem.2009.11.011. Beyersmann, D.; Hartwig, A. Carcinogenic metal compounds: Recent insight into molecular and cellular mechanisms. Arch. Toxicol. 2008, 82, 493–512. https://doi.org/10.1007/s00204-008-0313-y. (In English) Int. J. Environ. Res. Public Health 2022, 19, 4669 32. 33. 34. 35. 36. 37. 38. 39. 40. 41. 42. 43. 44. 45. 46. 47. 48. 49. 50. 51. 52. 53. 54. 55. 56. 57. 58. 59. 60. 61. 62. 63. 17 of 30 Ungureanu, G.; Santos, S.; Boaventura, R.; Botelho, C. Arsenic and antimony in water and wastewater: Overview of removal techniques with special reference to latest advances in adsorption. J. Environ. Manag. 2015, 151, 326–342. https://doi.org/10.1016/j.jenvman.2014.12.051. (In English) de la Calle-Guntiñas, M.B.; Madrid, Y.; Cámara, C. Stability study of total antimony, Sb(III) and Sb(V) at the trace level. Fresenius’ J. Anal. Chem. 1992, 344, 27–29. https://doi.org/10.1007/bf00324836. Zheng, J.; Ohata, M.; Furuta, N. Antimony Speciation in Environmental Samples by Using High-Performance Liquid Chromatography Coupled to Inductively Coupled Plasma Mass Spectrometry. Anal. Sci. 2000, 16, 75–80. https://doi.org/10.2116/analsci.16.75. Ho, T.-L. Hard soft acids bases (HSAB) principle and organic chemistry. Chem. Rev. 1975, 75, 1–20. https://doi.org/10.1021/cr60293a001. Burford, N.; Carpenter, Y.-Y.; Conrad, E.; Saunders, C.D.L. ChemInform Abstract: The Chemistry of Arsenic, Antimony and Bismuth. ChemInform 2012, 43, 1–17. https://doi.org/10.1002/chin.201246227. Wang, C.Y. Antimony: Its History, Chemistry, Mineralogy, Geology, Metallurgy, Uses, Preparations, Analysis, Production, and Valuation; with Complete Bibliographies. For Students, Manufacturers, and Users of Antimony; Griffin: Brussels, Belgium, 1909. Boyle, R.; Jonasson, I. The geochemistry of antimony and its use as an indicator element in geochemical prospecting. J. Geochem. Explor. 1984, 20, 223–302. https://doi.org/10.1016/0375-6742(84)90071-2. Kyono, A.; Kimata, M.; Matsuhisa, M.; Miyashita, Y.; Okamoto, K. Low-temperature crystal structures of stibnite implying orbital overlap of Sb 5s 2 inert pair electrons. Phys. Chem. Miner. 2002, 29, 254–260. https://doi.org/10.1007/s00269-001-0227-1. Kuze, S.; Saiki, A.; Du Boulay, D.; Ishizawa, N.; Pring, A. X-ray diffraction evidence for a monoclinic form of stibnite, Sb 2 S 3 , below 290 K. Am. Miner. 2004, 89, 1022–1025. https://doi.org/10.2138/am-2004-0712. Chang, L.L.Y.; Li, X.; Zheng, C. The Jamesonite-Benavidesite Series. Can. Mineral. 1987, 25, 667–672. And, Y.M.; Ueda, Y. Structure and Physical Properties of 1D Magnetic Chalcogenide, Jamesonite (FePb4Sb6S14). Inorg. Chem. 2003, 42, 7830–7838. https://doi.org/10.1021/ic034634t. Schaller, W.T. Crystallography of valentinite (Sb2O3) and andorite(?) (2PbS·Ag2S·3Sb2S3) from Oregon. Am. Mineral. 1937, 22, 651–666. Svensson, C. The crystal structure of orthorhombic antimony trioxide, Sb2O3. Acta Crystallogr. Sect. B 1974, 30, 458–461. https://doi.org/10.1107/s0567740874002986. Svensson, C. Refinement of the crystal structure of cubic antimony trioxide, Sb2O3. Acta Crystallogr. Sect. B 1975, 31, 2016–2018. https://doi.org/10.1107/s0567740875006759. Whitten, A.E.; Dittrich, B.; Spackman, M.A.; Turner, P.; Brown, T.C. Charge density analysis of two polymorphs of antimony(iii) oxide. Dalton Trans. 2004, 1, 23–29. https://doi.org/10.1039/b312550e. Vitaliano, C.J.; Mason, B. Stibiconite and cervantite. Am. Mineral. 1952, 37, 982–999. Christy, A.; Atencio, D. Clarification of status of species in the pyrochlore supergroup. Miner. Mag. 2013, 77, 13–20. https://doi.org/10.1180/minmag.2013.077.1.02. Bothwell, D.I.; Davis, R.J.; Moss, A.A. A Bismuth-Bearing Variety of Bindheimite. Mineral. Mag. J. Mineral. Soc. 1960, 32, 664–666. Cervelle, B. Détermination par microréflectométrie de propriétés optiques d’un cristal monoclinique absorbant (kermesite Sb2S2O). Deuxième partie. Bull. Minéralogie 1972, 95, 464–469. Baumgardt, E.; Kupcik, V. Synthesis of kermesite Sb2S2O. J. Cryst. Growth 1977, 37, 346–348. https://doi.org/10.1016/00220248(77)90130-0. Kharbish, S.; Libowitzky, E.; Beran, A. Raman spectra of isolated and interconnected pyramidal XS3 groups (X = Sb,Bi) in stibnite, bismuthinite, kermesite, stephanite and bournonite. Eur. J. Miner. 2009, 21, 325–333. https://doi.org/10.1127/09351221/2009/0021-1914. Tatsuka, K.; Morimoto, N. Tetrahedrite stability relations in the Cu-Fe-Sb-S system. Am. Mineral. 1977, 62, 1101–1109. Johnson, N.E.; Craig, J.R.; Rimstidt, J.D. Compositional trends in tetrahedrite. Can. Mineral. 1986, 24, 385–397. Wenk, H.-R.; Bulakh, A. Minerals: Their Constitution and Origin; Cambridge University Press: Cambridge, UK, 2004. Seal, R.R.I.; Schulz, K.J.; DeYoung, J.H.J. Antimony. In Critical Mineral Resources of the United States—Economic and Environmental Geology and Prospects for Future Supply; No. U.S. Geological Survey Professional Paper 1802; Schulz, K.J., DeYoung, J.H.J., Seal, R.R.I., Bradley, D.C., Eds.; U.S. Geological Survey: Washington, DC, USA, 2017; Volume 1802, pp. C1–C17. Miller, M.H. Antimony. In United States Mineral Resources; Brobst, D.A., Pratt, W.P., Eds.; U.S. Geological Survey Professional Paper: Washington, DC, USA, 1973; Volume 820, pp. 45–50. Pohl, W.L. Economic Geology Principles and Practice: Metals, Minerals, Coal and Hydrocarbons—Introduction to Formation and Sustainable Exploitation of Mineral Deposits; Wiley-Blackwell: Hoboken, NJ, USA, 2011. Schwarz-Schampera, U. Antimony (Critical Metals Handbook); Wiley and Sons: Hoboken, NJ, USA, 2014. Jiada, W. Antimony vein deposits of China. Ore Geol. Rev. 1993, 8, 213–232. https://doi.org/10.1016/0169-1368(93)90017-s. Williams-Jones, A.E.; Norman, C. Controls of mineral parageneses in the system Fe-Sb-S-O. Econ. Geol. 1997, 92, 308–324. https://doi.org/10.2113/gsecongeo.92.3.308. Diemar, G.A.; Filella, M.; Leverett, P.; Williams, P.A. Dispersion of antimony from oxidizing ore deposits. Pure Appl. Chem. 2009, 81, 1547–1553. https://doi.org/10.1351/pac-con-08-10-21. Pavlova, G.G.; Borisenko, A.S. The age of Ag–Sb deposits of Central Asia and their correlation with other types of ore systems and magmatism. Ore Geol. Rev. 2009, 35, 164–185. https://doi.org/10.1016/j.oregeorev.2008.11.006. Int. J. Environ. Res. Public Health 2022, 19, 4669 64. 65. 66. 67. 68. 69. 70. 71. 72. 73. 74. 75. 76. 77. 78. 79. 80. 81. 82. 83. 84. 85. 86. 87. 18 of 30 Pavlova, G.G.; Borovikov, A.A. Physicochemical factors of formation of Au-As, Au-Sb, and Ag-Sb deposits. Geol. Ore Deposits 2009, 50, 433–444. https://doi.org/10.1134/s1075701508060020. Bortnikov, N.S.; Gamynin, G.N.; Vikent’Eva, O.V.; Prokof’Ev, V.Y.; Prokop’Ev, A.V. The Sarylakh and Sentachan gold-antimony deposits, Sakha-Yakutia: A case of combined mesothermal gold-quartz and epithermal stibnite ores. Geol. Ore Deposits 2010, 52, 339–372. https://doi.org/10.1134/s1075701510050028. Fornadel, A.P.; Spry, P.G.; Melfos, V.; Vavelidis, M.; Voudouris, P.C. Is the Palea Kavala Bi–Te–Pb–Sb±Au district, northeastern Greece, an intrusion-related system? Ore Geol. Rev. 2011, 39, 119–133. https://doi.org/10.1016/j.oregeorev.2011.01.002. Melfos, V.; Voudouris, P.C. Geological, Mineralogical and Geochemical Aspects for Critical and Rare Metals in Greece. Minerals 2012, 2, 300–317. https://doi.org/10.3390/min2040300. Wang, Z.; Xia, Y.; Song, X.; Liu, J.; Yang, C.; Yan, B. Study on the evolution of ore-formation fluids for Au-Sb ore deposits and the mechanism of Au-Sb paragenesis and differentiation in the southwestern part of Guizhou Province, China. Chin. J. Geochem. 2013, 32, 56–68. https://doi.org/10.1007/s11631-013-0607-5. Melfos, V.; Voudouris, P. Cenozoic metallogeny of Greece and potential for precious, critical and rare metals exploration. Ore Geol. Rev. 2017, 89, 1030–1057. https://doi.org/10.1016/j.oregeorev.2017.05.029. Voudouris, P.; Spry, P.G.; Melfos, V.; Alfieris, D.; Mavrogonatos, C.; Repstock, A.; Djiba, A.; Stergiou, C.; Periferakis, A.; Melfou, M. Porphyry and Epithermal Deposits in Greece: A Review and New Discoveries. In Proceedings of the 8. Geochemistry Symposium, Antalya, Turkey, 2–6 May 2018; p. 181. Němec, M.; Zachariáš, J. The Krásná Hora, Milešov, and Příčovy Sb-Au ore deposits, Bohemian Massif: Mineralogy, fluid inclusions, and stable isotope constraints on the deposit formation. Miner. Depos. 2018, 53, 225–244. https://doi.org/10.1007/s00126-017-0734-8. Qiu, K.-F.; Yu, H.-C.; Deng, J.; McIntire, D.; Gou, Z.-Y.; Geng, J.-Z.; Chang, Z.-S.; Zhu, R.; Li, K.-N.; Goldfarb, R. The giant Zaozigou Au-Sb deposit in West Qinling, China: Magmatic- or metamorphic-hydrothermal origin? Miner. Depos. 2020, 55, 345– 362. https://doi.org/10.1007/s00126-019-00937-w. Hofstra, A.H.; Marsh, E.E.; Todorov, T.I.; Emsbo, P. Fluid inclusion evidence for a genetic link between simple antimony veins and giant silver veins in the Coeur d’Alene mining district, ID and MT, USA. Geofluids 2013, 13, 475–493. https://doi.org/10.1111/gfl.12036. WHO. Guidelines for Drinking-Water Quality; World Health Organization: Geneva, Switzerland, 2011; Volume 216, pp. 303–304. Lai, Z.; He, M.; Lin, C.; Ouyang, W.; Liu, X. Interactions of antimony with biomolecules and its effects on human health. Ecotoxicol. Environ. Saf. 2022, 233, 113317. https://doi.org/10.1016/j.ecoenv.2022.113317. Johnson, C.C.; Demetriades, A. Urban Geochemical Mapping: A Review of Case Studies in this Volume. In Mapping the Chemical Environment of Urban Areas; Johnson, C.C., Demetriades, A., Locutura, J., Ottesen, R.T., Eds.; Wiley: Hoboken, NJ, USA, 2011; pp. 7–27. Yesilonis, I.; Pouyat, R.; Neerchal, N. Spatial distribution of metals in soils in Baltimore, Maryland: Role of native parent material, proximity to major roads, housing age and screening guidelines. Environ. Pollut. 2008, 156, 723–731. https://doi.org/10.1016/j.envpol.2008.06.010. Guéguen, F.; Stille, P.; Geagea, M.L.; Boutin, R. Atmospheric pollution in an urban environment by tree bark biomonitoring– Part I: Trace element analysis. Chemosphere 2012, 86, 1013–1019. https://doi.org/10.1016/j.chemosphere.2011.11.040. Levresse, G.; Lopez, G.; Tritlla, J.; López, E.C.; Chavez, A.C.; Salvador, E.M.; Soler, A.; Corbella, M.; Sandoval, L.H.; CoronaEsquivel, R. Phytoavailability of antimony and heavy metals in arid regions: The case of the Wadley Sb district (San Luis, Potosí, Mexico). Sci. Total Environ. 2012, 427–428, 115–125. https://doi.org/10.1016/j.scitotenv.2012.04.020. Ramsey, M.H.; Argyraki, A. Estimation of measurement uncertainty from field sampling: Implications for the classification of contaminated land. Sci. Total Environ. 1997, 198, 243–257. https://doi.org/10.1016/s0048-9697(97)05456-9. Patinha, C.; Armienta, A.; Argyraki, A.; Durães, N. Chapter 6—Inorganic Pollutants in Soils. In Soil Pollution; Duarte, A.C., Cachada, A., Rocha-Santos, T., Eds.; Academic Press: Cambridge, MA, USA; 2018, pp. 127–159. Manta, D.S.; Angelone, M.; Bellanca, A.; Neri, R.; Sprovieri, M. Heavy metals in urban soils: A case study from the city of Palermo (Sicily), Italy. Sci. Total Environ. 2002, 300, 229–243. https://doi.org/10.1016/s0048-9697(02)00273-5. Rodrigues, S.; Urquhart, G.; Hossack, I.; Pereira, E.; Duarte, A.; Davidson, C.; Hursthouse, A.; Tucker, P.; Roberston, D. The influence of anthropogenic and natural geochemical factors on urban soil quality variability: A comparison between Glasgow, UK and Aveiro, Portugal. Environ. Chem. Lett. 2009, 7, 141–148. https://doi.org/10.1007/s10311-008-0149-y. Kelepertsis, A.; Argyraki, A.; Alexakis, D. Multivariate statistics and spatial interpretation of geochemical data for assessing soil contamination by potentially toxic elements in the mining area of Stratoni, north Greece. Geochem. Explor. Environ. Anal. 2006, 6, 349–355. https://doi.org/10.1144/1467-7873/05-101. Frenzel, M.; Voudouris, P.; Cook, N.J.; Ciobanu, C.L.; Gilbert, S.; Wade, B.P. Evolution of a hydrothermal ore-forming system recorded by sulfide mineral chemistry: A case study from the Plaka Pb–Zn–Ag Deposit, Lavrion, Greece. Miner. Depos. 2021, 57, 417–438. https://doi.org/10.1007/s00126-021-01067-y. Periferakis, A.; Paresoglou, N. Lavrion from Ancient Greece to the Present Day: A Study of how an Ore Deposit Shaped History. In Proceedings of the 15th International Congress of the Geological Society of Greece, Athens, Greece, 22–24 May 2019; pp. 704–705. Ross, J.; Voudouris, P.; Melfos, V.; Vaxevanopoulos, M.; Soukis, K.; Merigot, K. The Lavrion silver district: Reassessing its ancient mining history. Geoarchaeology 2021, 36, 617–642. https://doi.org/10.1002/gea.21852. Int. J. Environ. Res. Public Health 2022, 19, 4669 88. 89. 90. 91. 92. 93. 94. 95. 96. 97. 98. 99. 100. 101. 102. 103. 104. 105. 106. 107. 108. 109. 110. 111. 112. 19 of 30 Periferakis, A.; Paresoglou, I.; Paresoglou, N. The significance of the Lavrion mines in Greek and European Geoheritage. Eur. Geol. 2019, 48, 24–27. Papastamatiou, D.; Skarpelis, N.; Argyraki, A. Air Quality in Mining Areas: The Case of Stratoni, Chalkidiki, Greece. Bull. Geol. Soc. Greece 2017, 43, 2510–2519. https://doi.org/10.12681/bgsg.11657. Nocete, F.; Álex, E.; Nieto, J.M.; Sáez, R.; Rodríguez-Bayona, M. An archaeological approach to regional environmental pollution in the south-western Iberian Peninsula related to Third millennium BC mining and metallurgy. J. Archaeol. Sci. 2005, 32, 1566– 1576. https://doi.org/10.1016/j.jas.2005.04.012. Ferrier, G. Application of Imaging Spectrometer Data in Identifying Environmental Pollution Caused by Mining at Rodaquilar, Spain. Remote Sens. Environ. 1999, 68, 125–137. https://doi.org/10.1016/s0034-4257(98)00105-9. Vaseashta, A.; Vaclavikova, M.; Gallios, G.; Roy, P.; Pummakarnchana, O. Nanostructures in environmental pollution detection, monitoring, and remediation. Sci. Technol. Adv. Mater. 2007, 8, 47–59. https://doi.org/10.1016/j.stam.2006.11.003. Silva, L.; de Vallejuelo, S.F.O.; Martinez-Arkarazo, I.; Castro, K.; Oliveira, M.; Sampaio, C.H.; de Brum, I.A.; de Leão, F.B.; Taffarel, S.R.; Madariaga, J.M. Study of environmental pollution and mineralogical characterization of sediment rivers from Brazilian coal mining acid drainage. Sci. Total Environ. 2013, 447, 169–178. https://doi.org/10.1016/j.scitotenv.2012.12.013. Yurkevich, N.V.; Abrosimova, N.A.; Bortnikova, S.B.; Karin, Y.G.; Saeva, O.P. Geophysical investigations for evaluation of environmental pollution in a mine tailings area. Toxicol. Environ. Chem. 2017, 99, 1328–1345. https://doi.org/10.1080/02772248.2017.1371308. Periferakis, A. The Yukon Gold Rush: Early Examples of the Socioeconomic and Environmental Impact of Mining. In Proceedings of the 15th International Congress of the Geological Society of Greece, Athens, Greece, 22–24 May 2019; pp. 710–711. Adriano, D.C. Trace Elements in the Terrestrial Environment; Springer: Berlin, Germany, 1986. Telford, K.; Maher, W.; Krikowa, F.; Foster, S.; Ellwood, M.J.; Ashley, P.M.; Lockwood, P.V.; Wilson, S.C. Bioaccumulation of antimony and arsenic in a highly contaminated stream adjacent to the Hillgrove Mine, NSW, Australia. Environ. Chem. 2009, 6, 133–143. https://doi.org/10.1071/en08097. Wilson, N.; Webster-Brown, J. The fate of antimony in a major lowland river system, the Waikato River, New Zealand. Appl. Geochem. 2009, 24, 2283–2292. https://doi.org/10.1016/j.apgeochem.2009.09.016. Ragaini, R.C.; Ralston, H.R.; Roberts, N. Environmental trace metal contamination in Kellogg, Idaho, near a lead smelting complex. Environ. Sci. Technol. 1977, 11, 773–781. https://doi.org/10.1021/es60131a004. Ainsworth, N.; Cooke, J.A.; Johnson, M.S. Biological significance of antimony in contaminated grassland. Water Air Soil Pollut. 1991, 57, 193–199. https://doi.org/10.1007/bf00282882. Baroni, F.; Boscagli, A.; Protano, G.; Riccobono, F. Antimony accumulation in Achillea ageratum, Plantago lanceolata and Silene vulgaris growing in an old Sb-mining area. Environ. Pollut. 2000, 109, 347–352. https://doi.org/10.1016/s0269-7491(99)00240-7. Pacyna, J.M.; Pacyna, E.G. An assessment of global and regional emissions of trace metals to the atmosphere from anthropogenic sources worldwide. Environ. Rev. 2001, 9, 269–298. https://doi.org/10.1139/a01-012. Qi, C.; Liu, G.; Chou, C.-L.; Zheng, L. Environmental geochemistry of antimony in Chinese coals. Sci. Total Environ. 2008, 389, 225–234. https://doi.org/10.1016/j.scitotenv.2007.09.007. (In English) Tian, H.Z.; Zhao, D.; He, M.C.; Wang, Y.; Cheng, K. Temporal and spatial distribution of atmospheric antimony emission inventories from coal combustion in China. Environ. Pollut. 2011, 159, 1613–1619. https://doi.org/10.1016/j.envpol.2011.02.048. (In English) He, M.; Wang, N.; Long, X.; Zhang, C.; Ma, C.; Zhong, Q.; Wang, A.; Wang, Y.; Pervaiz, A.; Shan, J. Antimony speciation in the environment: Recent advances in understanding the biogeochemical processes and ecological effects. J. Environ. Sci. 2019, 75, 14–39. https://doi.org/10.1016/j.jes.2018.05.023. He, M.; Wang, X.; Wu, F.; Fu, Z. Antimony pollution in China. Sci. Total Environ. 2012, 421–422, 41–50. https://doi.org/10.1016/j.scitotenv.2011.06.009. Fu, Z.; Wu, F.; Amarasiriwardena, D.; Mo, C.; Liu, B.; Zhu, J.; Deng, Q.; Liao, H. Antimony, arsenic and mercury in the aquatic environment and fish in a large antimony mining area in Hunan, China. Sci. Total Environ. 2010, 408, 3403–3410. https://doi.org/10.1016/j.scitotenv.2010.04.031. Courtin-Nomade, A.; Rakotoarisoa, O.; Bril, H.; Grybos, M.; Forestier, L.; Foucher, F.; Kunz, M. Weathering of Sb-rich mining and smelting residues: Insight in solid speciation and soil bacteria toxicity. Geochemistry 2012, 72, 29–39. https://doi.org/10.1016/j.chemer.2012.02.004. Hiller, E.; Lalinská, B.; Chovan, M.; Jurkovič, L.; Klimko, T.; Jankulár, M.; Hovorič, R.; Šottník, P.; Fľaková, R.; Ženišová, Z.; et al. Arsenic and antimony contamination of waters, stream sediments and soils in the vicinity of abandoned antimony mines in the Western Carpathians, Slovakia. Appl. Geochem. 2012, 27, 598–614. https://doi.org/10.1016/j.apgeochem.2011.12.005. Cidu, R.; Biddau, R.; Dore, E.; Vacca, A.; Marini, L. Antimony in the soil–water–plant system at the Su Suergiu abandoned mine (Sardinia, Italy): Strategies to mitigate contamination. Sci. Total Environ. 2014, 497–498, 319–331. https://doi.org/10.1016/j.scitotenv.2014.07.117. (In English) Macgregor, K.; MacKinnon, G.; Farmer, J.G.; Graham, M.C. Mobility of antimony, arsenic and lead at a former antimony mine, Glendinning, Scotland. Sci. Total Environ. 2015, 529, 213–222. https://doi.org/10.1016/j.scitotenv.2015.04.039. Wilson, N.; Craw, D.; Hunter, K. Antimony distribution and environmental mobility at an historic antimony smelter site, New Zealand. Environ. Pollut. 2004, 129, 257–266. https://doi.org/10.1016/j.envpol.2003.10.014. Int. J. Environ. Res. Public Health 2022, 19, 4669 20 of 30 113. Mykolenko, S.; Liedienov, V.; Kharytonov, M.; Makieieva, N.; Kuliush, T.; Queralt, I.; Marguí, E.; Hidalgo, M.; Pardini, G.; Gispert, M. Presence, mobility and bioavailability of toxic metal(oids) in soil, vegetation and water around a Pb-Sb recycling factory (Barcelona, Spain). Environ. Pollut. 2018, 237, 569–580. https://doi.org/10.1016/j.envpol.2018.02.053. 114. Casado, M.; Anawar, H.M.; Garcia-Sanchez, A.; Regina, I.S. Antimony and Arsenic Uptake by Plants in an Abandoned Mining Area. Commun. Soil Sci. Plant Anal. 2007, 38, 1255–1275. https://doi.org/10.1080/00103620701328412. 115. Periferakis, A. The Keramos Antimonite Mines in Chios Island, Greece: Mining History and Current Situation. News Miner. 2020, 35, 5–21. 116. Karczewska, A.; Bogda, A.; Krysiak, A. Arsenic in soils in the areas of former mining and mineral processing in Lower Silesia, southwestern Poland. In Trace Metals and Other Contaminants in the Environment; Elsevier: Amsterdam, The Netherlands, 2007; Volume 9, pp. 411–440. 117. Karczewska, A.; Krysiak, A.; Mokrzycka, D.; Jezierski, P.; Szopka, K. Arsenic Distribution in Soils of a Former As Mining Area and Processing. Pol. J. Environ. Stud. 2013, 22, 175–181. Available online: http://www.pjoes.com/Arsenic-Distribution-in-Soilsof-a-Former-As-r-nMining-Area-and-Processing,88966,0,2.html (accessed on 13 January 2022). 118. Lewińska, K.; Karczewska, A. Antimony in soils of SW Poland—an overview of potentially enriched sites. Environ. Monit. Assess. 2019, 191, 70. https://doi.org/10.1007/s10661-019-7214-9. (In English) 119. Chatzidiakos, E.; Fanouraki, M.; Kelepertsis, A.; Argyraki, A.; Alexakis, D. Speciation and mobility of Arsenic and Antimony in groundwater at Melivoia, East Thessaly and Keramos area NW Chios, Greece. In Proceedings of the 8th International Hydrogeological Congress of Greece, Athens, Greece, 8–10 October 2008; Volume 1, pp. 219–228. 120. McCallum, R.I. Occupational exposure to antimony compounds. J. Environ. Monit. 2005, 7, 1245–1250. https://doi.org/10.1039/b509118g. 121. Elmaaboud, R.M.A.; Mohamed, Z.T.; George, S.M.; El-Dine, A.M.E.; El Shehaby, D.M. Lead and Cadmium Toxicity in Tile Manufacturing Workers in Assiut, Egypt. Arab J. Forensic Sci. Forensic Med. 2016, 1, 299–311. https://doi.org/10.12816/0026462. 122. Bienert, G.P.; Thorsen, M.; Schüssler, M.D.; Nilsson, H.R.; Wagner, A.; Tamás, M.J.; Jahn, T.P. A subgroup of plant aquaporins facilitate the bi-directional diffusion of As(OH)3 and Sb(OH)3across membranes. BMC Biol. 2008, 6, 26. https://doi.org/10.1186/1741-7007-6-26. 123. Kamiya, T.; Fujiwara, T. Arabidopsis NIP1;1 Transports Antimonite and Determines Antimonite Sensitivity. Plant Cell Physiol. 2009, 50, 1977–1981. https://doi.org/10.1093/pcp/pcp130. 124. Baes, C.F.; Mesmer, R.S. The Hydrolysis of Cations; Berichte der Bunsengesellschaft für physikalische Chemie, No. 2; John Wiley & Sons: New York, NY, USA, 1977; p. 489. 125. Filella, M.; Belzile, N.; Lett, M.-C. Antimony in the environment: A review focused on natural waters. III. Microbiota relevant interactions. Earth-Sci. Rev. 2007, 80, 195–217. https://doi.org/10.1016/j.earscirev.2006.09.003. 126. Pokrovski, G.S.; Borisova, A.Y.; Roux, J.; Hazemann, J.-L.; Petdang, A.; Tella, M.; Testemale, D. Antimony speciation in saline hydrothermal fluids: A combined X-ray absorption fine structure spectroscopy and solubility study. Geochim. Cosmochim. Acta 2006, 70, 4196–4214. https://doi.org/10.1016/j.gca.2006.06.1549. 127. Mitsunobu, S.; Harada, T.; Takahashi, Y. Comparison of Antimony Behavior with that of Arsenic under Various Soil Redox Conditions. Environ. Sci. Technol. 2006, 40, 7270–7276. https://doi.org/10.1021/es060694x. (In English) 128. Scheinost, A.C.; Rossberg, A.; Vantelon, D.; Xifra, I.; Kretzschmar, R.; Leuz, A.-K.; Funke, H.; Johnson, C.A. Quantitative antimony speciation in shooting-range soils by EXAFS spectroscopy. Geochim. Cosmochim. Acta 2006, 70, 3299–3312. https://doi.org/10.1016/j.gca.2006.03.020. 129. Oorts, K.; Smolders, E.; Degryse, F.; Buekers, J.; Gascó, G.; Cornelis, G.; Mertens, J. Solubility and Toxicity of Antimony Trioxide (Sb2O3) in Soil. Environ. Sci. Technol. 2008, 42, 4378–4383. https://doi.org/10.1021/es703061t. (In English) 130. Deng, T.; Chen, Y.-W.; Belzile, N. Antimony speciation at ultra trace levels using hydride generation atomic fluorescence spectrometry and 8-hydroxyquinoline as an efficient masking agent. Anal. Chim. Acta 2001, 432, 293–302. https://doi.org/10.1016/s0003-2670(00)01387-8. 131. Chen, Y.W.; Deng, T.L.; Filella, M.; Belzile, N. Distribution and Early Diagenesis of Antimony Species in Sediments and Porewaters of Freshwater Lakes. Environ. Sci. Technol. 2003, 37, 1163–1168. https://doi.org/10.1021/es025931k. (In English) 132. Brooks, R.R. Geobotany and Biogeochemestry in Mineral Exploration; Harper & Row: New York, NY, USA, 1972. 133. Bowen, H.J.M. Environmental Chemistry of the Elements; Academic Press: London, UK, 1979. 134. Jung, M.C.; Thornton, I.; Chon, H.-T. Arsenic, Sb and Bi contamination of soils, plants, waters and sediments in the vicinity of the Dalsung Cu–W mine in Korea. Sci. Total Environ. 2002, 295, 81–89. https://doi.org/10.1016/s0048-9697(02)00042-6. 135. De Gregori, I.; Fuentes, E.; Rojas, M.; Pinochet, H.; Potin-Gautier, M. Monitoring of copper, arsenic and antimony levels in agricultural soils impacted and non-impacted by mining activities, from three regions in Chile. J. Environ. Monit. 2003, 5, 287– 295. https://doi.org/10.1039/b211469k. (In English) 136. Miravet, R.; Bonilla, E.; López-Sánchez, J.F.; Rubio, R. Antimony speciation in terrestrial plants. Comparative studies on extraction methods. J. Environ. Monit. 2005, 7, 1207–1213. https://doi.org/10.1039/b509115b. 137. Tschan, M.; Robinson, B.; Schulin, R. Antimony uptake by Zea mays (L.) and Helianthus annuus (L.) from nutrient solution. Environ. Geochem. Health 2008, 30, 187–191. https://doi.org/10.1007/s10653-008-9142-4. 138. Fu, Z.; Wu, F.; Mo, C.; Liu, B.; Zhu, J.; Deng, Q.; Liao, H.; Zhang, Y. Bioaccumulation of antimony, arsenic, and mercury in the vicinities of a large antimony mine, China. Microchem. J. 2011, 97, 12–19. https://doi.org/10.1016/j.microc.2010.06.004. Int. J. Environ. Res. Public Health 2022, 19, 4669 21 of 30 139. Hammel, W.; Debus, R.; Steubing, L. Mobility of antimony in soil and its availability to plants. Chemosphere 2000, 41, 1791–1798. https://doi.org/10.1016/s0045-6535(00)00037-0. (In English) 140. Argyraki, A.; Kelepertzis, E. Urban soil geochemistry in Athens, Greece: The importance of local geology in controlling the distribution of potentially harmful trace elements. Sci. Total Environ. 2014, 482–483, 366–377. https://doi.org/10.1016/j.scitotenv.2014.02.133. 141. Slooff, W.; Bont, P.F.H.; Hesse, J.M.; Loos, B. Exploratory report Antimony and antimony compounds. In Scopingsrapport Antimoon En Antimoonverbindingen; National Institute of Public Health and Environmental Protection: Bilthoven, The Netherlands, 1992. 142. Salma, I.; Maenhaut, W. Changes in elemental composition and mass of atmospheric aerosol pollution between 1996 and 2002 in a Central European city. Environ. Pollut. 2006, 143, 479–488. https://doi.org/10.1016/j.envpol.2005.11.042. 143. Gaman, L.; Delia, C.E.; Luzardo, O.P.; Zumbado, M.; Badea, M.; Stoian, I.; Gilca, M.; Boada, L.D.; Henríquez-Hernández, L.A. Serum concentration of toxic metals and rare earth elements in children and adolescent. Int. J. Environ. Health Res. 2020, 30, 696– 712. https://doi.org/10.1080/09603123.2019.1626353. 144. Belzile, N.; Chen, Y.-W.; Filella, M. Human Exposure to Antimony: I. Sources and Intake. Crit. Rev. Environ. Sci. Technol. 2011, 41, 1309–1373. https://doi.org/10.1080/10643381003608227. 145. Fu, F.; Wang, Q. Removal of heavy metal ions from wastewaters: A review. J. Environ. Manag. 2011, 92, 407–418. https://doi.org/10.1016/j.jenvman.2010.11.011. (In English) 146. Kang, M.; Kamei, T.; Magara, Y. Comparing polyaluminum chloride and ferric chloride for antimony removal. Water Res. 2003, 37, 4171–4179. https://doi.org/10.1016/s0043-1354(03)00351-8. (In English) 147. Guo, X.; Wu, Z.; He, M. Removal of antimony(V) and antimony(III) from drinking water by coagulation–flocculation–sedimentation (CFS). Water Res. 2009, 43, 4327–4335. https://doi.org/10.1016/j.watres.2009.06.033. (In English) 148. Daneshvar, E.; Vazirzadeh, A.; Niazi, A.; Kousha, M.; Naushad, M.; Bhatnagar, A. Desorption of Methylene blue dye from brown macroalga: Effects of operating parameters, isotherm study and kinetic modeling. J. Clean. Prod. 2017, 152, 443–453. https://doi.org/10.1016/j.jclepro.2017.03.119. 149. Albadarin, A.B.; Collins, M.N.; Naushad, M.; Shirazian, S.; Walker, G.; Mangwandi, C. Activated lignin-chitosan extruded blends for efficient adsorption of methylene blue. Chem. Eng. J. 2017, 307, 264–272. https://doi.org/10.1016/j.cej.2016.08.089. 150. Inam, M.A.; Khan, R.; Park, D.R.; Khan, S.; Uddin, A.; Yeom, I.T. Complexation of Antimony with Natural Organic Matter: Performance Evaluation during Coagulation-Flocculation Process. Int. J. Environ. Res. Public Health 2019, 16, 1092. https://doi.org/10.3390/ijerph16071092. (In English) 151. Wu, Z.; He, M.; Guo, X.; Zhou, R. Removal of antimony (III) and antimony (V) from drinking water by ferric chloride coagulation: Competing ion effect and the mechanism analysis. Sep. Purif. Technol. 2010, 76, 184–190. https://doi.org/10.1016/j.seppur.2010.10.006. 152. Tang, X.; Zheng, H.; Teng, H.; Sun, Y.; Guo, J.; Xie, W.; Yang, Q.; Chen, W. Chemical coagulation process for the removal of heavy metals from water: A review. Desalination Water Treat. 2016, 57, 1733–1748. https://doi.org/10.1080/19443994.2014.977959. 153. Buschmann, J.; Sigg, L. Antimony(III) Binding to Humic Substances: Influence of pH and Type of Humic Acid. Environ. Sci. Technol. 2004, 38, 4535–4541. https://doi.org/10.1021/es049901o. (In English) 154. Filella, M.; Williams, P.A.; Belzile, N. Antimony in the environment: Knowns and unknowns. Environ. Chem. 2009, 6, 95–105. https://doi.org/10.1071/en09007. 155. Shotyk, W.; Krachler, M.; Chen, B. Contamination of Canadian and European bottled waters with antimony from PET containers. J. Environ. Monit. 2006, 8, 288–292. https://doi.org/10.1039/b517844b. (In English) 156. A Ward, L.; Cain, O.L.; A Mullally, R.; Holliday, K.S.; Wernham, A.G.; Baillie, P.D.; Greenfield, S.M. Health beliefs about bottled water: A qualitative study. BMC Public Health 2009, 9, 196. https://doi.org/10.1186/1471-2458-9-196. 157. Hu, Z.; Morton, L.W.; Mahler, R.L. Bottled Water: United States Consumers and Their Perceptions of Water Quality. Int. J. Environ. Res. Public Health 2011, 8, 565–578. https://doi.org/10.3390/ijerph8020565. (In English) 158. Sevigny, C. The Success of Bottled Water: The Hidden Costs Hurt Us and the Environment. Bachelor’s Thesis, University of Montana, Missoula, MT, USA, 2017. 159. Qian, N. Bottled Water or Tap Water? A Comparative Study of Drinking Water Choices on University Campuses. Water 2018, 10, 59. Available online: https://www.mdpi.com/2073-4441/10/1/59 (accessed on 13 January 2022). 160. Vieux, F.; Maillot, M.; Rehm, C.D.; Barrios, P.L.; Drewnowski, A. Trends in tap and bottled water consumption among children and adults in the United States: Analyses of NHANES 2011-16 data. Nutr. J. 2020, 19, 10. https://doi.org/10.1186/s12937-0200523-6. 161. Shotyk, W.; Krachler, M. Contamination of Bottled Waters with Antimony Leaching from Polyethylene Terephthalate (PET) Increases upon Storage. Environ. Sci. Technol. 2007, 41, 1560–1563. https://doi.org/10.1021/es061511+. 162. Westerhoff, P.; Prapaipong, P.; Shock, E.; Hillaireau, A. Antimony leaching from polyethylene terephthalate (PET) plastic used for bottled drinking water. Water Res. 2008, 42, 551–556. https://doi.org/10.1016/j.watres.2007.07.048. 163. Keresztes, S.; Tatár, E.; Mihucz, V.; Virág, I.; Majdik, C.; Záray, G. Leaching of antimony from polyethylene terephthalate (PET) bottles into mineral water. Sci. Total Environ. 2009, 407, 4731–4735. https://doi.org/10.1016/j.scitotenv.2009.04.025. 164. Kalač, P.; Svoboda, L.R. A review of trace element concentrations in edible mushrooms. Food Chem. 2000, 69, 273–281. https://doi.org/10.1016/s0308-8146(99)00264-2. 165. Borovička, J.; Řanda, Z.; Jelínek, E. Antimony content of macrofungi from clean and polluted areas. Chemosphere 2006, 64, 1837– 1844. https://doi.org/10.1016/j.chemosphere.2006.01.060. Int. J. Environ. Res. Public Health 2022, 19, 4669 22 of 30 166. He, M. Distribution and phytoavailability of antimony at an antimony mining and smelting area, Hunan, China. Environ. Geochem. Health 2007, 29, 209–219. https://doi.org/10.1007/s10653-006-9066-9. (In English) 167. Cava-Montesinos, P.; de la Guardia, A.; Teutsch, C.; Cervera, M.L.; de la Guardia, M. Non-chromatographic speciation analysis of arsenic and antimony in milk hydride generation atomic fluorescence spectrometry. Anal. Chim. Acta 2003, 493, 195–203. https://doi.org/10.1016/s0003-2670(03)00876-6. 168. Cava-Montesinos, P. Determination of arsenic and antimony in milk by hydride generation atomic fluorescence spectrometry. Talanta 2003, 60, 787–799. https://doi.org/10.1016/s0039-9140(03)00126-7. (In English) 169. Waheed, S.; Zaidi, J.H.; Ahmad, S. Instrumental neutron activation analysis of 23 individual food articles from a high altitude region. J. Radioanal. Nucl. Chem. Artic. 2003, 258, 73–81. https://doi.org/10.1023/a:1026253924758. 170. Lund, W. Determination of arsenic and antimony in wine by electrothermal atomic absorption spectrometry. Anal. Bioanal. Chem. 1996, 354, 93–96. https://doi.org/10.1007/s002169600015. 171. Garg, A.N.; Ramakrishna, V.V.S. Fish as an indicator of aquatic environment: Multielemental neutron activation analysis of nutrient and pollutant elements in fish from Indian coastal areas. Toxicol. Environ. Chem. 2006, 88, 125–140. https://doi.org/10.1080/02772240500241793. 172. Hansen, H.R.; Pergantis, S.A. Detection of antimony species in citrus juices and drinking water stored in PET containers. J. Anal. At. Spectrom. 2006, 21, 731–733. https://doi.org/10.1039/b606367e. 173. Zheng, J.; Iijima, A.; Furuta, N. Complexation effect of antimony compounds with citric acid and its application to the speciation of antimony(iii) and antimony(v) using HPLC-ICP-MS. J. Anal. At. Spectrom. 2001, 16, 812–818. https://doi.org/10.1039/b101943k. 174. Khlifi, R.; Hamza-Chaffai, A. Head and neck cancer due to heavy metal exposure via tobacco smoking and professional exposure: A review. Toxicol. Appl. Pharmacol. 2010, 248, 71–88. https://doi.org/10.1016/j.taap.2010.08.003. (In English) 175. Ibanez, Y.; Le Bot, B.; Glorennec, P. House-dust metal content and bioaccessibility: A review. Eur. J. Miner. 2010, 22, 629–637. https://doi.org/10.1127/0935-1221/2010/0022-2010. 176. Wiseman, C.L. Analytical methods for assessing metal bioaccessibility in airborne particulate matter: A scoping review. Anal. Chim. Acta 2015, 877, 9–18. https://doi.org/10.1016/j.aca.2015.01.024. 177. Pelfrêne, A.; Cave, M.r.; Wragg, J.; Douay, F. In Vitro Investigations of Human Bioaccessibility from Reference Materials Using Simulated Lung Fluids. Int. J. Environ. Res. Public Health 2017, 14, 112. Available online: https://www.mdpi.com/16604601/14/2/112 (accessed on 13 January 2022). 178. Kelepertzis, E.; Chrastný, V.; Botsou, F.; Sigala, E.; Kypritidou, Z.; Komárek, M.; Skordas, K.; Argyraki, A. Tracing the sources of bioaccessible metal(loid)s in urban environments: A multidisciplinary approach. Sci. Total Environ. 2021, 771, 144827. https://doi.org/10.1016/j.scitotenv.2020.144827. 179. Allen, J.P.; Carey, J.J.; Walsh, A.; Scanlon, D.O.; Watson, G.W. Electronic Structures of Antimony Oxides. J. Phys. Chem. C 2013, 117, 14759–14769. https://doi.org/10.1021/jp4026249. 180. Smichowski, P.; Madrid, Y.; Guntiñas, M.B.D.L.C.; Cámara, C. Separation and determination of antimony(III) and antimony(V) species by high-performance liquid chromatography with hydride generation atomic absorption spectrometric and inductively coupled plasma mass spectrometric detection. J. Anal. At. Spectrom. 1995, 10, 815–821. https://doi.org/10.1039/JA9951000815. 181. Delnomdedieu, M.; Basti, M.M.; Otvos, J.D.; Thomas, D.J. Reduction and binding of arsenate and dimethylarsinate by glutathione: A magnetic resonance study. Chem. Biol. Interact. 1994, 90, 139–155. https://doi.org/10.1016/0009-2797(94)90099-x. (In English) 182. Tirmenstein, M.; Mathias, P.; Snawder, J.; Wey, H.; Toraason, M. Antimony-induced alterations in thiol homeostasis and adenine nucleotide status in cultured cardiac myocytes. Toxicology 1997, 119, 203–211. https://doi.org/10.1016/s0300-483x(97)03628-7. 183. Gebel, T. Arsenic and antimony: Comparative approach on mechanistic toxicology. Chem. Interact. 1997, 107, 131–144. https://doi.org/10.1016/s0009-2797(97)00087-2. 184. Grover, P.; Rekhadevi, P.; Danadevi, K.; Vuyyuri, S.; Mahboob, M.; Rahman, M. Genotoxicity evaluation in workers occupationally exposed to lead. Int. J. Hyg. Environ. Health 2010, 213, 99–106. https://doi.org/10.1016/j.ijheh.2010.01.005. (In English) 185. García-Lestón, J.; Roma-Torres, J.; Vilares, A.M.; Pinto, R.M.; Prista, J.; Teixeira, J.P.; Mayan, O.; Conde, J.; Pingarilho, M.; Gaspar, J.; et al. Genotoxic effects of occupational exposure to lead and influence of polymorphisms in genes involved in lead toxicokinetics and in DNA repair. Environ. Int. 2012, 43, 29–36. https://doi.org/10.1016/j.envint.2012.03.001. (In English) 186. Bocca, B.; Pino, A.; Alimonti, A.; Forte, G. Toxic metals contained in cosmetics: A status report. Regul. Toxicol. Pharmacol. 2014, 68, 447–467. https://doi.org/10.1016/j.yrtph.2014.02.003. (In English) 187. El Shanawany, S.; Foda, N.; Hashad, D.I.; Salama, N.; Sobh, Z. The potential DNA toxic changes among workers exposed to antimony trioxide. Environ. Sci. Pollut. Res. 2017, 24, 12455–12461. https://doi.org/10.1007/s11356-017-8805-z. (In English) 188. Hayat, F.; Shah, S.N.A.; Rehman, Z.U.; Bélanger-Gariepy, F. Antimony(III) dithiocarbamates: Crystal structures, supramolecular aggregations, DNA binding, antioxidant and antileishmanial activities. Polyhedron 2021, 194, 114909. https://doi.org/10.1016/j.poly.2020.114909. 189. Asghar, F.; Badshah, A.; Shah, A.; Rauf, M.K.; Ali, M.I.; Tahir, M.N.; Nosheen, E.; Rehman, Z.U.; Qureshi, R. Synthesis, characterization and DNA binding studies of organoantimony(V) ferrocenyl benzoates. J. Organomet. Chem. 2012, 717, 1–8. https://doi.org/10.1016/j.jorganchem.2012.07.028. 190. Cavallo, D.; Iavicoli, I.; Setini, A.; Marinaccio, A.; Perniconi, B.; Carelli, G.; Iavicoli, S. Genotoxic risk and oxidative DNA damage in workers exposed to antimony trioxide. Environ. Mol. Mutagen. 2002, 40, 184–189. https://doi.org/10.1002/em.10102. (In English) Int. J. Environ. Res. Public Health 2022, 19, 4669 23 of 30 191. Kirkland, D.; Whitwell, J.; Deyo, J.; Serex, T. Failure of antimony trioxide to induce micronuclei or chromosomal aberrations in rat bone-marrow after sub-chronic oral dosing. Mutat. Res. 2007, 627, 119–128. https://doi.org/10.1016/j.mrgentox.2006.10.012. (In English) 192. Hashemzaei, M.; Pourahmad, J.; Safaeinejad, F.; Tabrizian, K.; Akbari, F.; Bagheri, G.; Hosseini, M.-J.; Shahraki, J. Antimony induces oxidative stress and cell death in normal hepatocytes. Toxicol. Environ. Chem. 2015, 97, 256–265. https://doi.org/10.1080/02772248.2015.1032616. 193. Seiple, L.A.; Cardellina, J.H., 2nd; Akee, R.; Stivers, J.T. Potent Inhibition of Human Apurinic/Apyrimidinic Endonuclease 1 by Arylstibonic Acids. Mol. Pharmacol. 2007, 73, 669–677. https://doi.org/10.1124/mol.107.042622. (In English) 194. Phillips, M.A.; Cánovas, A.; Wu, P.-W.; Islas-Trejo, A.; Medrano, J.F.; Rice, R.H. Parallel responses of human epidermal keratinocytes to inorganic SbIII and AsIII. Environ. Chem. 2016, 13, 963–970. https://doi.org/10.1071/EN16019. (In English) 195. Morales, M.E.; Derbes, R.S.; Ade, C.M.; Ortego, J.C.; Stark, J.; Deininger, P.L.; Roy-Engel, A.M. Heavy Metal Exposure Influences Double Strand Break DNA Repair Outcomes. PLoS ONE 2016, 11, e0151367. https://doi.org/10.1371/journal.pone.0151367. (In English) 196. Jiang, X.; An, Z.; Lu, C.; Chen, Y.; Du, E.; Qi, S.; Yang, K.; Zhang, Z.; Xu, Y. The protective role of Nrf2-Gadd45b against antimony-induced oxidative stress and apoptosis in HEK293 cells. Toxicol. Lett. 2016, 256, 11–18. https://doi.org/10.1016/j.toxlet.2016.05.016. (In English) 197. Zhang, D.; Lee, D.-J.; Pan, X. Desorption of Hg(II) and Sb(V) on extracellular polymeric substances: Effects of pH, EDTA, Ca(II) and temperature shocks. Bioresour. Technol. 2013, 128, 711–715. https://doi.org/10.1016/j.biortech.2012.10.089. 198. Xiaojian, L.; Xingkang, J.; Ming, G.; Yousheng, K.; Dongliang, P.; Ningchen, L.; Sijin, L. Non-toxic Dose of Antimony Exposure Could Enhance the Intracellular Energy Metabolism and Promote Prostate Cancer Progression. Asian J. Ecotoxicol. 2015, 10, 129–135. 199. Wu, C.; Li, F.; Xu, H.; Zeng, W.; Yu, R.; Wu, X.; Shen, L.; Liu, Y.; Li, J. The potential role of brassinosteroids (BRs) in alleviating antimony (Sb) stress in Arabidopsis thaliana. Plant Physiol. Biochem. 2019, 141, 51–59. https://doi.org/10.1016/j.plaphy.2019.05.011. 200. Xia, S.; Zhu, X.; Yan, Y.; Zhang, T.; Chen, G.; Lei, D.; Wang, G. Developmental neurotoxicity of antimony (Sb) in the early life stages of zebrafish. Ecotoxicol. Environ. Saf. 2021, 218, 112308. https://doi.org/10.1016/j.ecoenv.2021.112308. 201. Park, G.; Brock, D.J.; Pellois, J.-P.; Gabbaï, F.P. Heavy Pnictogenium Cations as Transmembrane Anion Transporters in Vesicles and Erythrocytes. Chem 2019, 5, 2215–2227. https://doi.org/10.1016/j.chempr.2019.06.013. 202. Huang, X.; Zhang, B.; Wu, L.; Zhou, Y.; Li, Y.; Mao, X.; Chen, Y.; Wang, J.; Luo, P.; Ma, J.; et al. Association of Exposure to Ambient Fine Particulate Matter Constituents With Semen Quality Among Men Attending a Fertility Center in China. Environ. Sci. Technol. 2019, 53, 5957–5965. https://doi.org/10.1021/acs.est.8b06942. 203. Zafar, A.; Eqani, S.A.M.A.S.; Bostan, N.; Cincinelli, A.; Tahir, F.; Shah, S.T.A.; Hussain, A.; Alamdar, A.; Huang, Q.; Peng, S.; et al. Toxic metals signature in the human seminal plasma of Pakistani population and their potential role in male infertility. Environ. Geochem. Health 2015, 37, 515–527. https://doi.org/10.1007/s10653-014-9666-8. 204. Zhang, G.; Wang, X.; Zhang, X.; Li, Q.; Xu, S.; Huang, L.; Zhang, Y.; Lin, L.; Gao, D.; Wu, M.; et al. Antimony in urine during early pregnancy correlates with increased risk of gestational diabetes mellitus: A prospective cohort study. Environ. Int. 2019, 123, 164–170. https://doi.org/10.1016/j.envint.2018.11.072. 205. Zhang, Q.; Li, X.; Liu, X.; Dong, M.; Xiao, J.; Wang, J.; Zhou, M.; Wang, Y.; Ning, D.; Ma, W.; et al. Association between maternal antimony exposure and risk of gestational diabetes mellitus: A birth cohort study. Chemosphere 2020, 246, 125732. https://doi.org/10.1016/j.chemosphere.2019.125732. 206. Vigeh, M.; Yunesian, M.; Matsukawa, T.; Shamsipour, M.; Jeddi, M.Z.; Rastkari, N.; Hassanvand, M.S.; Shariat, M.; Kashani, H.; Pirjani, R.; et al. Prenatal blood levels of some toxic metals and the risk of spontaneous abortion. J. Environ. Health Sci. Eng. 2021, 19, 357–363. https://doi.org/10.1007/s40201-020-00608-3. 207. Bienert, G.P.; Schüssler, M.D.; Jahn, T.P. Metalloids: Essential, beneficial or toxic? Major intrinsic proteins sort it out. Trends Biochem. Sci. 2008, 33, 20–26. https://doi.org/10.1016/j.tibs.2007.10.004. (In English) 208. Rosen, B.P.; Tamás, M.J. Arsenic Transport in Prokaryotes and Eukaryotic Microbes. Adv. Exp. Med. Biol. 2010, 679, 47–55. (In English) 209. Mukhopadhyay, R.; Bhattacharjee, H.; Rosen, B.P. Aquaglyceroporins: Generalized metalloid channels. Biochim. Biophys. Acta BBA Gen. Subj. 2014, 1840, 1583–1591. https://doi.org/10.1016/j.bbagen.2013.11.021. (In English) 210. Sanders, O.I.; Rensing, C.; Kuroda, M.; Mitra, B.; Rosen, B.P. Antimonite is accumulated by the glycerol facilitator GlpF in Escherichia coli. J. Bacteriol. 1997, 179, 3365–3367. https://doi.org/10.1128/jb.179.10.3365-3367.1997. 211. Meng, Y.L.; Liu, Z.; Rosen, B.P. As(III) and Sb(III) Uptake by GlpF and Efflux by ArsB in Escherichia coli. J. Biol. Chem. 2004, 279, 18334–18341. https://doi.org/10.1074/jbc.m400037200. (In English) 212. Hachez, C.; Chaumont, F. Aquaporins: A Family of Highly Regulated Multifunctional Channels. In MIPs and Their Role in the Exchange of Metalloids; Advances in Experimental Medicine and Biology, No. 679; Jahn, T.P., Bienert, G.P., Eds.; Springer: New York, NY, USA, 2010; pp. 1–17. 213. Wysocki, R.; Chery, C.C.; Wawrzycka, D.; Van Hulle, M.; Cornelis, R.; Thevelein, J.; Tamas, M.J. The glycerol channel Fps1p mediates the uptake of arsenite and antimonite in Saccharomyces cerevisiae. Mol. Microbiol. 2001, 40, 1391–1401. https://doi.org/10.1046/j.1365-2958.2001.02485.x. (In English) 214. Liu, Z.; Shen, J.; Carbrey, J.M.; Mukhopadhyay, R.; Agre, P.; Rosen, B.P. Arsenite transport by mammalian aquaglyceroporins AQP7 and AQP9. Proc. Natl. Acad. Sci. USA 2002, 99, 6053–6058. https://doi.org/10.1073/pnas.092131899. (In English) Int. J. Environ. Res. Public Health 2022, 19, 4669 24 of 30 215. Liu, Z.; Carbrey, J.M.; Agre, P.; Rosen, B.P. Arsenic trioxide uptake by human and rat aquaglyceroporins. Biochem. Biophys. Res. Commun. 2004, 316, 1178–1185. https://doi.org/10.1016/j.bbrc.2004.03.003. (In English) 216. Yang, H.-C.; Cheng, J.; Finan, T.M.; Rosen, B.P.; Bhattacharjee, H. Novel Pathway for Arsenic Detoxification in the Legume Symbiont Sinorhizobium meliloti. J. Bacteriol. 2005, 187, 6991–6997. https://doi.org/10.1128/jb.187.20.6991-6997.2005. (In English) 217. Maciaszczyk-Dziubinska, E.; Migdal, I.; Migocka, M.; Bocer, T.; Wysocki, R. The yeast aquaglyceroporin Fps1p is a bidirectional arsenite channel. FEBS Lett. 2009, 584, 726–732. https://doi.org/10.1016/j.febslet.2009.12.027. (In English) 218. Liu, Z.; Sanchez, M.A.; Jiang, X.; Boles, E.; Landfear, S.; Rosen, B.P. Mammalian glucose permease GLUT1 facilitates transport of arsenic trioxide and methylarsonous acid. Biochem. Biophys. Res. Commun. 2006, 351, 424–430. https://doi.org/10.1016/j.bbrc.2006.10.054. (In English) 219. Liu, Z.; Boles, E.; Rosen, B.P. Arsenic Trioxide Uptake by Hexose Permeases in Saccharomyces cerevisiae. J. Biol. Chem. 2004, 279, 17312–17318. https://doi.org/10.1074/jbc.m314006200. (In English) 220. Maciaszczyk-Dziubinska, E.; Wawrzycka, D.; Wysocki, R. Arsenic and Antimony Transporters in Eukaryotes. Int. J. Mol. Sci. 2012, 13, 3527–3548. https://doi.org/10.3390/ijms13033527. (In English) 221. Zangi, R.; Filella, M. Transport routes of metalloids into and out of the cell: A review of the current knowledge. Chem. Interact. 2012, 197, 47–57. https://doi.org/10.1016/j.cbi.2012.02.001. (In English) 222. Frézard, F.; Demicheli, C.; Ferreira, C.S.; Costa, M.A.P. Glutathione-Induced Conversion of Pentavalent Antimony to Trivalent Antimony in Meglumine Antimoniate. Antimicrob. Agents Chemother. 2001, 45, 913–916. https://doi.org/10.1128/aac.45.3.913916.2001. (In English) 223. Yan, S.; Wong, I.L.K.; Chow, L.M.C.; Sun, H. Rapid reduction of pentavalent antimony by trypanothione: Potential relevance to antimonial activation. Chem. Commun. 2003, 266–267. https://doi.org/10.1039/b210240d. 224. Yan, S.; Li, F.; Ding, K.; Sun, H. Reduction of pentavalent antimony by trypanothione and formation of a binary and ternary complex of antimony(III) and trypanothione. JBIC J. Biol. Inorg. Chem. 2003, 8, 689–697. https://doi.org/10.1007/s00775-003-04681. (In English) 225. Denton, H.; McGregor, J.C.; Coombs, G.H. Reduction of anti-leishmanial pentavalent antimonial drugs by a parasite-specific thiol-dependent reductase, TDR1. Biochem. J. 2004, 381, 405–412. https://doi.org/10.1042/bj20040283. (In English) 226. Cobbett, C.; Goldsbrough, P. Phytochelatins and Metallothioneins: Roles in heavy Metal Detoxification and Homeostasis. Annu. Rev. Plant Biol. 2002, 53, 159–182. https://doi.org/10.1146/annurev.arplant.53.100301.135154. (In English) 227. Verbruggen, N.; Hermans, C.; Schat, H. Mechanisms to cope with arsenic or cadmium excess in plants. Curr. Opin. Plant Biol. 2009, 12, 364–372. https://doi.org/10.1016/j.pbi.2009.05.001. (In English) 228. Wysocki, R.; Tamás, M.J. HowSaccharomyces cerevisiaecopes with toxic metals and metalloids. FEMS Microbiol. Rev. 2010, 34, 925–951. https://doi.org/10.1111/j.1574-6976.2010.00217.x. (In English) 229. Lemire, J.A.; Harrison, J.J.; Turner, R.J. Antimicrobial activity of metals: Mechanisms, molecular targets and applications. Nat. Rev. Microbiol. 2013, 11, 371–384. https://doi.org/10.1038/nrmicro3028. 230. Ge, R.; Sun, H. Bioinorganic Chemistry of Bismuth and Antimony: Target Sites of Metallodrugs. Acc. Chem. Res. 2007, 40, 267– 274. https://doi.org/10.1021/ar600001b. (In English) 231. Adeyemi, J.O.; Onwudiwe, D.C. Chemistry and Some Biological Potential of Bismuth and Antimony Dithiocarbamate Complexes. Molecules 2020, 25, 305. https://doi.org/10.3390/molecules25020305. (In English) 232. Scott, N.; Hatlelid, K.M.; MacKenzie, N.E.; Carter, D.E. Reactions of arsenic(III) and arsenic(V) species with glutathione. Chem. Res. Toxicol. 1993, 6, 102–106. https://doi.org/10.1021/tx00031a016. (In English) 233. Sun, H.; Yan, S.C.; Cheng, W.S. Interaction of antimony tartrate with the tripeptide glutathione. JBIC J. Biol. Inorg. Chem. 2000, 267, 5450–5457. https://doi.org/10.1046/j.1432-1327.2000.01605.x. (In English) 234. Pitman, A.L.; Pourbaix, M.; De Zoubov, N. Potential-pH Diagram of the Antimony-Water System: Its Applications to Properties of the Metal, Its Compounds, Its Corrosion, and Antimony Electrodes. J. Electrochem. Soc. 1957, 104, 594. 235. Kip, A.E.; Schellens, J.H.M.; Beijnen, J.H.; Dorlo, T.P.C. Clinical Pharmacokinetics of Systemically Administered Antileishmanial Drugs. Clin. Pharmacokinet. 2017, 57, 151–176. https://doi.org/10.1007/s40262-017-0570-0. 236. Wang, Y.-X.; Pan, A.; Feng, W.; Liu, C.; Huang, L.-L.; Ai, S.-H.; Zeng, Q.; Lu, W.-Q. Variability and exposure classification of urinary levels of non-essential metals aluminum, antimony, barium, thallium, tungsten and uranium in healthy adult men. J. Expo. Sci. Environ. Epidemiol. 2019, 29, 424–434. https://doi.org/10.1038/s41370-017-0002-0. (In English) 237. Barregard, L.; Ellingsen, D.G.; Berlinger, B.; Weinbruch, S.; Harari, F.; Sallsten, G. Normal variability of 22 elements in 24-h urine samples–Results from a biobank from healthy non-smoking adults. Int. J. Hyg. Environ. Health 2021, 233, 113693. 238. Ye, L.; Qiu, S.; Li, X.; Jiang, Y.; Jing, C. Antimony exposure and speciation in human biomarkers near an active mining area in Hunan, China. Sci. Total Environ. 2018, 640–641, 1–8. https://doi.org/10.1016/j.scitotenv.2018.05.267. 239. Diaz-Bone, R.A.; Van de Wiele, T. Biotransformation of metal(loid)s by intestinal microorganisms. Pure Appl. Chem. 2010, 82, 409–427. https://doi.org/10.1351/pac-con-09-06-08. 240. Patriarca, M.; Menditto, A.; Rossi, B.; Lyon, T.; Fell, G. Environmental exposure to metals of newborns, infants and young children. Microchem. J. 2000, 67, 351–361. https://doi.org/10.1016/s0026-265x(00)00088-6. 241. Brieger, H.; Semisch, C.W., 3rd; Stasney, J.; Piatnek, D.A. Industrial antimony poisoning. Ind. Med. Surg. 1954, 23, 521–523. (In English) 242. Klucik, I.; Kemka, L.R. The excretion of antimony in workers in antimony metallurgical works (Czech.). Prac. Lek. 1960, 12, 133–138. Int. J. Environ. Res. Public Health 2022, 19, 4669 25 of 30 243. McCallum, R.I. The Work of an Occupational Hygiene Service in Environmental Control. Ann. Occup. Hyg. 1963, 6, 55–64,. https://doi.org/10.1093/annhyg/6.2.55. 244. Cooper, D.A.; Pendergrass, E.P.; Vorwald, A.J.; Mayock, R.L.; Brieger, H. Pneumoconiosis among workers in an Antimony Industry. Am. J. Roentgenol. 1968, 103, 495–508. https://doi.org/10.2214/ajr.103.3.495. 245. Lüdersdorf, R.; Fuchs, A.; Mayer, P.; Skulsuksai, G.; Schäcke, G. Biological assessment of exposure to antimony and lead in the glass-producing industry. Int. Arch. Occup. Environ. Health 1987, 59, 469–474. https://doi.org/10.1007/bf00377841. (In English) 246. Bailly, R.; Lauwerys, R.; Buchet, J.P.; Mahieu, P.; Konings, J. Experimental and human studies on antimony metabolism: Their relevance for the biological monitoring of workers exposed to inorganic antimony. Occup. Environ. Med. 1991, 48, 93–97. https://doi.org/10.1136/oem.48.2.93. (In English) 247. Belyaeva, A.P. The effect of antimony on reproduction. Gig. Tr. Prof Zabol. 1967, 11, 32–37. 248. McCallum, R. Detection of Antimony in Process Workers’ Lungs by X-Radiation. Occup. Med. 1967, 17, 134–138. https://doi.org/10.1093/occmed/17.4.134. (In English) 249. McCallum, R.I.; Day, M.J.; Underhill, J.; Aird, E.G. Measurement of antimony oxide dust in human lungs in vivo by X-ray spectrophotometry. Inhaled Part. 1970, 2, 611–619. (In English) 250. Gerhardsson, L.; Brune, D.; Nordberg, G.F.; Wester, P.O. Antimony in lung, liver and kidney tissue from deceased smelter workers. Scand. J. Work. Environ. Health 1982, 8, 201–208. https://doi.org/10.5271/sjweh.2475. (In English) 251. Vanoeteren, C.; Cornelis, R.; Versieck, J. Evaluation of trace elements in human lung tissue I. Concentration and distribution. Sci. Total Environ. 1986, 54, 217–230. https://doi.org/10.1016/0048-9697(86)90267-6. 252. Dernehl, C.U.; Nau, C.A.; Sweets, H.H. Animal studies on the toxicity of inhaled antimony trioxide. J. Ind. Hyg. Toxicol. 1945, 27, 256–262. (In English) 253. Bulmer, F.M.R.; Johnston, J.H. Antimony trisulfide. J. Ind. Hyg. Toxicol. 1948, 30, 26–28. (In English) 254. Gross, P.; Westrick, M.L.; Brown, J.H.; Srsic, R.P.; Schrenk, H.H.; Hatch, T.F. Toxicologic study of calcium halophosphate phosphors and antimony trioxide. II. Pulmonary studies. AMA Arch. Ind. Health 1955, 11, 479–486. (In English) 255. Felicetti, S.A.; Thomas, R.G.; McClellan, R.O. Metabolism of Two Valence States of Inhaled Antimony in Hamsters. AIHAJ 1974, 35, 292–300. https://doi.org/10.1080/0002889748507037. 256. Leffler, P.; Gerhardsson, L.; Brune, D.; Nordberg, G.F. Lung retention of antimony and arsenic in hamsters after the intratracheal instillation of industrial dust. Scand. J. Work. Environ. Health 1984, 10, 245–251. https://doi.org/10.5271/sjweh.2330. (In English) 257. Groth, D.H.; Stettler, L.E.; Burg, J.R.; Busey, W.M.; Grant, G.C.; Wong, L. Carcinogenic effects of antimony trioxide and antimony ore concentrate in rats. J. Toxicol. Environ. Health Part A 1986, 18, 607–626. https://doi.org/10.1080/15287398609530898. 258. Newton, P.E.; Bolte, H.F.; Daly, I.W.; Pillsbury, B.D.; Terrill, J.B.; Drew, R.T.; Ben-Dyke, R.; Sheldon, A.W.; Rubin, L.F. Subchronic and Chronic Inhalation Toxicity of Antimony Trioxide in the Rat. Fundam. Appl. Toxicol. 1994, 22, 561–576. https://doi.org/10.1006/faat.1994.1063. (In English) 259. Taylor, P.J. Acute Intoxication from Antimony Trichloride. Occup. Environ. Med. 1966, 23, 318–321. https://doi.org/10.1136/oem.23.4.318. (In English) 260. Renes, L.E. Antimony poisoning in industry. AMA Arch. Ind. Hyg. Occup. Med. 1953, 7, 99–108. (In English) 261. Klucik, I.; Juck, A.; Gruberova, J. Respiratory and pulmonary lesions caused by antimony trioxide dust. Prac. Lek. 1962, 14, 363–368. 262. McCallum, R.I. The industrial toxicology of antimony. The Ernestine Henry lecture 1987. J. R. Coll. Physicians Lond. 1989, 23, 28– 32. (In English) 263. Bradley, W.R.; Fredrick, W.G. The Toxicity of Antimony:—Animal Studies—. Am. Ind. Hyg. Assoc. Q. 1941, 2, 15–22. https://doi.org/10.1080/00968204109343800. 264. Honey, M. The effects of sodium antimony tartrate on the myocardium. Br. Heart J. 1960, 22, 601–616. https://doi.org/10.1136/hrt.22.5.601. (In English) 265. Winship, K.A. Toxicity of antimony and its compounds. Advers. Drug React. Acute Poisoning Rev. 1987, 6, 67–90. (In English) 266. Hepburn, N.C.; Nolan, J.; Fenn, L.; Herd, R.M.; Neilson, J.M.; Sutherland, G.R.; Fox, K.A. Cardiac effects of sodium stibogluconate: Myocardial, electrophysiological and biochemical studies. QJM Int. J. Med. 1994, 87, 465–472. (In English) 267. Tirmenstein, M.; Plews, P.; Walker, C.; Woolery, M.; Wey, H.; Toraason, M. Antimony-Induced Oxidative Stress and Toxicity in Cultured Cardiac Myocytes. Toxicol. Appl. Pharmacol. 1995, 130, 41–47. https://doi.org/10.1006/taap.1995.1006. (In English) 268. Okamoto, Y.; Hidaka, S. Studies on calcium phosphate precipitation: Effects of metal ions used in dental materials. J. Biomed. Mater. Res. 1994, 28, 1403–1410. https://doi.org/10.1002/jbm.820281204. (In English) 269. Eke, C.; Er, K.; Segebade, C.; Boztosun, I. Study of filling material of dental composites: An analytical approach using radioactivation. Radiochim. Acta 2018, 106, 69–77. https://doi.org/10.1515/ract-2017-2766. 270. Imai, K.; Nakamura, M. In vitro embryotoxicity testing of metals for dental use by differentiation of embryonic stem cell test. Congenit. Anom. 2006, 46, 34–38. https://doi.org/10.1111/j.1741-4520.2006.00099.x. (In English) 271. Léonard, A.; Gerber, G. Mutagenicity, carcinogenicity and teratogenicity of antimony compounds. Mutat. Res. Genet. Toxicol. 1996, 366, 1–8. https://doi.org/10.1016/s0165-1110(96)90003-2. (In English) 272. Gebel, T. Suppression of arsenic-induced chromosome mutagenicity by antimony. Mutat. Res. Toxicol. Environ. Mutagen. 1998, 412, 213–218. https://doi.org/10.1016/s1383-5718(97)00181-2. (In English) 273. Davis, E.; Bakulski, K.M.; Goodrich, J.M.; Peterson, K.E.; Marazita, M.L.; Foxman, B. Low levels of salivary metals, oral microbiome composition and dental decay. Sci. Rep. 2020, 10, 14640. https://doi.org/10.1038/s41598-020-71495-9. (In English) Int. J. Environ. Res. Public Health 2022, 19, 4669 26 of 30 274. Holgerson, P.L.; Öhman, C.; Rönnlund, A.; Johansson, I. Maturation of Oral Microbiota in Children with or without Dental Caries. PLoS ONE 2015, 10, e0128534. https://doi.org/10.1371/journal.pone.0128534. (In English) 275. Caufield, P.; Schön, C.; Saraithong, P.; Li, Y.; Argimón, S. Oral Lactobacilli and Dental Caries: A Model for Niche Adaptation in Humans. J. Dent. Res. 2015, 94, 110S–118S. https://doi.org/10.1177/0022034515576052. (In English) 276. Wiener, R.C.; Bhandari, R. Association of electronic cigarette use with lead, cadmium, barium, and antimony body burden: NHANES 2015–2016. J. Trace Elem. Med. Biol. 2020, 62, 126602. https://doi.org/10.1016/j.jtemb.2020.126602. (In English) 277. Andrewes, P.; Cullen, W.R. Organoantimony Compounds in the Environment. In Organometallic Compounds in the Environment, 2nd ed.; Craig, P.G., Ed.; Wilry: Chichester, UK, 2003; pp. 277–303. 278. Falta, T.; Limbeck, A.; Koellensperger, G.; Hann, S. Bioaccessibility of selected trace metals in urban PM2.5 and PM10 samples: A model study. Anal. Bioanal. Chem. 2008, 390, 1149–1157. https://doi.org/10.1007/s00216-007-1762-5. (In English) 279. Dunn, J.T. A curious case of antimony poisoning Analyst 1928, 53, 532–533. 280. Lauwers, L.F.; Roelants, A.; Rosseel, P.M.; Heyndrickx, B.; Baute, L. Oral antimony intoxications in man. Crit. Care Med. 1990, 18, 324–326. https://doi.org/10.1097/00003246-199003000-00017. (In English) 281. Moskalev, Y.I. Materials on the distribution of radioactive antimony. Med. Radiol. 1959, 4, 6–13. 282. Waitz, J.A.; Ober, R.E.; Meisenhelder, J.E.; Thompson, P.E. Physiological disposition of antimony after administration of 124Sblabelled tartar emetic to rats, mice and monkeys, and the effects of tris (p-aminophenyl) Carbonium pamoate on this distribution. Bull. World Health Organ. 1965, 33, 537–546. (In English) 283. Van Bruwaene, R.; Gerber, G.B.; Kirchmann, R.; Colard, J. Metabolism of antimony-124 in lactating dairy cows. Health Phys. 1982, 43, 733–738. (In English) 284. Gerber, G.B.; Maes, J.; Eykens, B. Transfer of antimony and arsenic to the developing organism. Arch. Toxicol. 1982, 49, 159–168. https://doi.org/10.1007/bf00332363. (In English) 285. Dieter, M.P.; Jameson, C.W.; Elwell, M.R.; Lodge, J.W.; Hejtmancik, M.; Grumbein, S.L.; Ryan, M.; Peters, A.C. Comparative toxicity and tissue distribution of antimony potassium tartrate in rats and mice dosed by drinking water or intraperitoneal injection. J. Toxicol. Environ. Health Part A 1991, 34, 51–82. https://doi.org/10.1080/15287399109531548. (In English) 286. Subramanian, K.S.; Poon, R.; Chu, I.; Connor, J.W. Antimony in Drinking Water, Red Blood Cells, and Serum: Development of Analytical Methodology Using Transversely Heated Graphite Furnace Atomization-Atomic Absorption Spectrometry. Arch. Environ. Contam. Toxicol. 1997, 32, 431–435. https://doi.org/10.1007/s002449900209. (In English) 287. Gross, P.; Brown, J.H.; Westrick, M.L.; Srsic, R.P.; Butler, N.L.; Hatch, T.F. Toxicologic study of calcium halophosphate phosphors and antimony trioxide. I. Acute and chronic toxicity and some pharmacologic aspects. AMA Arch. Ind. Health 1955, 11, 473–478. (In English) 288. Schroeder, H.A.; Mitchener, M.; Nason, A.P. Zirconium, Niobium, Antimony, Vanadium and Lead in Rats: Life term studies. J. Nutr. 1970, 100, 59–68. https://doi.org/10.1093/jn/100.1.59. 289. Sunagawa, S. Experimental studies on antimony poisoning (author’s transl). Igaku kenkyu. 1981, 51, 129–142. (In Japanese). 290. Hiraoka, N. The toxicity and organ-distribution of antimony after chronic administration to rats. Kyoto Fenitsu Ika Daigaku Gasshi 1986, 95, 997–1017. 291. Rossi, F.; Acampora, R.; Vacca, C.; Maione, S.; Matera, M.G.; Servodio, R.; Marmo, E. Prenatal and postnatal antimony exposure in rats: Effect on vasomotor reactivity development of pups. Teratog. Carcinog. Mutagen. 1987, 7, 491–496. https://doi.org/10.1002/tcm.1770070507. (In English) 292. Poon, R.; Chu, I.; Lecavalier, P.; Valli, V.; Foster, W.; Gupta, S.; Thomas, B. Effects of antimony on rats following 90-day exposure via drinking water. Food Chem. Toxicol. 1998, 36, 21–35. https://doi.org/10.1016/s0278-6915(97)80120-2. (In English) 293. Oliver, T. The health of antimony oxide workers. BMJ 1933, 1, 1094–1095. https://doi.org/10.1136/bmj.1.3781.1094. 294. Stevenson, C.J. Antimony spots. Trans. St. John’s Hosp. Dermatol. Soc. 1965, 51, 40–48. (In English) 295. Thivolet, J.; Melinat, M.; Pellerat, J.; Perrot, H.; Francou, M. Occupational dermatitis attributed to antimony. Arch. Mal. Prof. 1971, 32, 571–573. (In French). 296. White Jr, G.P.; Mathias, C.G.; Davin, J.S. Dermatitis in workers exposed to antimony in a melting process. J. Occup. Med. 1993, 35, 392–395. (In English) 297. Volis, M.J. Dermatology Technique: Mohs Micrographic Surgery. Mako NSU Undergrad. Stud. J. 2021, 2021, 3. 298. Fukuyama, Y.; Kawarai, S.; Tezuka, T.; Kawabata, A.; Maruo, T. The palliative efficacy of modified Mohs paste for controlling canine and feline malignant skin wounds. Veter-Q. 2016, 36, 176–182. https://doi.org/10.1080/01652176.2015.1130880. 299. Mors, F.E.; Sevringhaus, E.L.; Schmidt, E.R. Conservative amputation of gangrenous parts by chemosurgery. Ann. Surg. 1941, 114, 274–282. https://doi.org/10.1097/00000658-194108000-00011. (In English) 300. Phelan, J.T. Mohs’ Chemosurgery Technique for Basal Cell Carcinoma of the Chin and Cheek Areas of the Face. Arch. Surg. 1963, 87, 212–214. https://doi.org/10.1001/archsurg.1963.01310140020007. 301. Lo, J.S.; Snow, S.N.; Mohs, F.E. Cylindroma Treated by Mohs Micrographic Surgery. J. Dermatol. Surg. Oncol. 1991, 17, 871–874. https://doi.org/10.1111/j.1524-4725.1991.tb03277.x. 302. Mohs, F.E. Mohs Micrographic Surgery: A Historical Perspective. Dermatol. Clin. 1989, 7, 609–612. https://doi.org/10.1016/s07338635(18)30556-4. 303. Trost, L.B.; Bailin, P.L. History of Mohs Surgery. Dermatol. Clin. 2011, 29, 135–139. https://doi.org/10.1016/j.det.2011.01.010. (In English) 304. Wang, M.Z.; Warshaw, E.M. Bloodroot. Derm. Clin. 2012, 23, 281–283. https://doi.org/10.1097/der.0b013e318273a4dd. Int. J. Environ. Res. Public Health 2022, 19, 4669 27 of 30 305. Mohs, F.E. Chemosurgery: A Microscopically Controlled Method of Cancer Excision. Arch. Surg. 1941, 42, 279–295. https://doi.org/10.1001/archsurg.1941.01210080079004. 306. Mohs, F.E.; Guyer, M.F. Pre-excisional fixation of tissues in the treatment of cancer in rats. Cancer Res. 1941, 1, 49–51. 307. Mohs, F.E. Chemosurgical treatment of cancer of the Lip: A Microscopically Controlled Method of Excision. Arch. Surg. 1944, 48, 478–488. https://doi.org/10.1001/archsurg.1944.01230010491009. 308. Mohs, F.E. Chemosurgical treatment of cancer of the nose: A microscopically controlled method. Arch. Surg. 1946, 53, 327–344. https://doi.org/10.1001/archsurg.1946.01230060334005. (In English) 309. Mohs, F.E. Chemosurgery for Melanoma. Arch. Dermatol. 1977, 113, 285. https://doi.org/10.1001/archderm.1977.01640030031003. 310. Mohs, F.E.; Snow, S.N.; Messing, E.M.; Kuglitsch, M.E. Microscopically Controlled Surgery in the Treatment of Carcinoma of the Penis. J. Urol. 1985, 133, 961–966. https://doi.org/10.1016/s0022-5347(17)49334-7. (In English) 311. Mohs, F.E. Micrographic surgery for the microscopically controlled excision of eyelid cancer: History and development. In Advances in Opthalmic Plastic and Reconstructive Surgery; Bosniak, S.L., Smith, B.C., Eds.; Pergamon Press: New York, NY, USA, 1986; pp. 381–408. 312. Mohs, F.; Larson, P.; Iriondo, M. Micrographic surgery for the microscopically controlled excision of carcinoma of the external ear. J. Am. Acad. Dermatol. 1988, 19, 729–737. https://doi.org/10.1016/s0190-9622(88)70229-7. (In English) 313. Croaker, A.; King, G.J.; Pyne, J.H.; Anoopkumar-Dukie, S.; Liu, L. Sanguinaria canadensis: Traditional Medicine, Phytochemical Composition, Biological Activities and Current Uses. Int. J. Mol. Sci. 2016, 17, 1414. Available online: https://www.mdpi.com/1422-0067/17/9/1414 (accessed on 13 January 2022). 314. Mohs, F.E. Chemosurgery for skin cancer: Fixed tissue and fresh tissue techniques. Arch. Dermatol. 1976, 112, 211–215. (In English) 315. Finley, E.M. The principles of mohs micrographic surgery for cutaneous neoplasia. Ochsner J. 2003, 5, 22–33. 316. Hobbs, E.R.; Wheeland, R.G.; Bailin, P.L.; Ratz, J.L.; Yetman, R.J.; Zins, J.E. Treatment of Dermatofibrosarcoma Protuberans with Mohs Micrographic Surgery. Ann. Surg. 1988, 207, 102–107. https://doi.org/10.1097/00000658-198801000-00019. (In English) 317. Komine, N.; Narita, S.; Kigure, T.; Tsuruta, H.; Numakura, K.; Akihama, S.; Saito, M.; Inoue, T.; Tsuchiya, N.; Satoh, S.; et al. Successful Local Control of Recurrent Penile Cancer Treated with a Combination of Systemic Chemotherapy, Irradiation, and Mohs’ Paste: A Case Report. Case Rep. Oncol. 2014, 7, 522–527. https://doi.org/10.1159/000365809. (In English) 318. Takeuchi, M.; Katsuki, T.; Yoshida, K.; Onoda, M.; Iwamura, M.; Inokuchi, T.; Furutani, A.; Katoh, T.; Kawano, K.; Hirata, K. Successful Pre-Operative Local Control of Skin Invasion of Breast Cancer Using a Combination of Systemic Chemotherapy and Mohs Paste. J. Breast Cancer 2021, 24, 481–490. https://doi.org/10.4048/jbc.2021.24.e46. (In English) 319. Firmino, F.; Villela-Castro, D.L.; dos Santos, J.; Santos, V.L.C.D.G. Topical Management of Bleeding From Malignant Wounds Caused by Breast Cancer: A Systematic Review. J. Pain Symptom Manag. 2021, 61, 1278–1286. https://doi.org/10.1016/j.jpainsymman.2020.10.020. 320. Yanazume, S.; Douzono, H.; Yanazume, Y.; Iio, K.; Douchi, T. New hemostatic method using Mohs’ paste for fatal genital bleeding in advanced cervical cancer. Gynecol. Oncol. Case Rep. 2013, 4, 47–49. https://doi.org/10.1016/j.gynor.2013.01.001. (In English) 321. Haldar, A.K.; Sen, P.; Roy, S. Use of Antimony in the Treatment of Leishmaniasis: Current Status and Future Directions. Mol. Biol. Int. 2011, 2011, 571242. https://doi.org/10.4061/2011/571242. 322. Fakhry, A. Asphyxia following injection of tartar emetic. Lancet 1931, 218, 1325. https://doi.org/10.1016/s0140-6736(00)99716-2. 323. Fry, W.B. Antimony in the treatment of syphilis. J. R. Army Med. Corps 1914, 22, 514–520. 324. Large, D.T.M.; Bonavia, V.J. Arsenic and Antimony in Malaria. J. R. Army Med. Corps 1926, 47, 430–438. https://doi.org/10.1136/jramc-47-06-04. 325. Patrick, A. Experiences with Intravenous Injections of Quinine and Antimony in the Treatment of Malaria. J. R. Army Med. Corps 1919, 32, 407–429. https://doi.org/10.1136/jramc-32-06-01. 326. Dye, W.H. Comparative Results in the Treatment of Frambœsia Tropica in Northern Nyasaland. J. R. Army Med. Corps 1924, 42, 280–286. https://doi.org/10.1136/jramc-42-04-05. 327. Ariza-Roldán, A.O.; López-Cardoso, E.M.; Rosas-Valdez, M.E.; Roman-Bravo, P.P.; Vargas-Pineda, D.G.; Cea-Olivares, R.; Acevedo-Quiroz, M.; Razo-Hernández, R.S.; Alvarez-Fitz, P.; Jancik, V. Synthesis, characterization, antimicrobial and theoretical studies of the first main group tris (ephedrinedithiocarbamate) complexes of As (III), Sb (III), Bi (III), Ga (III) and In (III). Polyhedron 2017, 134, 221–229. 328. Sharma, D.K.; Singh, Y.; Sharma, J. Monophenylantimony (III) Derivatives of Cyclic Dithiocarbamates; Synthesis, Spectroscopic Characterization, and Antimicrobial Study. Phosphorus Sulfur Silicon Relat. Elem. 2013, 188, 1194–1204. 329. Urgut, O.; Ozturk, I.; Banti, C.; Kourkoumelis, N.; Manoli, M.; Tasiopoulos, A.; Hadjikakou, S. New antimony(III) halide complexes with dithiocarbamate ligands derived from thiuram degradation: The effect of the molecule’s close contacts on in vitro cytotoxic activity. Mater. Sci. Eng. C 2016, 58, 396–408. https://doi.org/10.1016/j.msec.2015.08.030. (In English) 330. Duffin, J.; Campling, B.G. Therapy and disease concepts: The history (and future?) of antimony in cancer. J. Hist. Med. Allied Sci. 2002, 57, 61–78. https://doi.org/10.1093/jhmas/57.1.61. (In English) 331. Hunt, R.; McCANN, W.S.; Rowntree, L.G.; Voegtlin, C.; Eggleston, C.; Maxcy, K.F. The Status of Intravenous Therapy: V. Limitations to the use of quinine intravenously in the treatment of malaria. J. Am. Med Assoc. 1928, 91, 1372–1375. https://doi.org/10.1001/jama.1928.92700180003012. 332. Mitjà, O.; Hays, R.; Rinaldi, A.; McDermott, R.; Bassat, Q. New Treatment Schemes for Yaws: The Path Toward Eradication. Clin. Infect. Dis. 2012, 55, 406–412. https://doi.org/10.1093/cid/cis444. Int. J. Environ. Res. Public Health 2022, 19, 4669 28 of 30 333. Thakur, C.P.; Kumar, M.; Singh, S.K.; Sharma, D.; Prasad, U.S.; Singh, R.S.; Dhawan, P.S.; Achari, V. Comparison of regimens of treatment with sodium stibogluconate in kala-azar. BMJ 1984, 288, 895–897. https://doi.org/10.1136/bmj.288.6421.895. (In English) 334. Thakur, C.P.; Kumar, M.; Kumar, P.; Mishra, B.N.; Pandey, A.K. Rationalisation of regimens of treatment of kala-azar with sodium stibogluconate in India: A randomised study. BMJ 1988, 296, 1557–1561. https://doi.org/10.1136/bmj.296.6636.1557. (In English) 335. Chulay, J.D.; Mugambi, M.; Spencer, H.C. Electrocardiographic changes during Treatment of Leishmaniasis with Pentavalent Antimony (Sodium Stibogluconate). Am. J. Trop. Med. Hyg. 1985, 34, 702–709. https://doi.org/10.4269/ajtmh.1985.34.702. (In English) 336. Kouvoutsakis, G.; Mitsi, C.; Tarantilis, P.A.; Polissiou, M.G.; Pappas, C.S. Antimony compounds in the treatment of trypanosomiasis. Lancet 1910, 175, 938–939. https://doi.org/10.1016/s0140-6736(00)51330-0. 337. Jennings, F.W. Chemotherapy of trypanosomiasis: The potentiation of antimonial compounds by difluoromethylornithine (DFMO). Trop. Med. Parasitol. 1991, 42, 135–138. (In English) 338. Ercoli, N.; Minelli, E.B.; Olivo, N. Antitrypanosomal Activity of Trivalent Antimonials in vitro and Its Significance. Chemother. 1980, 26, 254–262. https://doi.org/10.1159/000237914. (In English) 339. Ercoli, N.; Minelli, E.B.; Villarroel, G. Chemotherapy of Trypanosoma venezuelense (T. evansi), I. Activity of trivalent antimonials in mice by long and short term tests. Ann. Trop. Med. Parasitol. 1980, 74, 485–493. (In English) 340. Christopherson, J. The Successful Use of Antimony in Bilharziosis: Administered as Intravenous Injections of Antimonium Tartaratum (Tartar Emetic). Lancet 1918, 192, 325–327. https://doi.org/10.1016/s0140-6736(01)02807-0. 341. Christopherson, J.B. Intravenous Injections of Antimony Tartrate in Bilharziasis. Lancet 1919, 194, 299. https://doi.org/10.1016/s0140-6736(01)44584-3. 342. Taylor, F.E. Intravenous Injections of Antimonium Tartaratum (Tartar Emetic) in Bilharziasis. J. R. Army Med. Corps 1919, 33, 181–190. https://doi.org/10.1136/jramc-33-02-09. 343. Cawston, F.G. The Use of Emetine in Treating Bilharzia Disease in the Child. J. R. Army Med. Corps 1926, 46, 57–60. https://doi.org/10.1136/jramc-46-01-08. 344. Alves, W.; Gelfand, M. Treatment of schistosomiasis with sodium antimony tri-gluconate by mouth. Trans. R. Soc. Trop. Med. Hyg. 1952, 46, 543–546. https://doi.org/10.1016/0035-9203(52)90046-1. 345. Davis, A. Comparative trials of antimonial drugs in urinary schistosomiasis. Bull. World Health Organ. 1968, 38, 197–227. 346. Pedrique, M.R.; Ercoli, N. Experimental and clinical studies with a new antimonial preparation for the treatment of schistosomiasis. Bull. World Health Organ. 1971, 45, 411–417. 347. Farid, Z.; Bassily, S.; Kent, D.C.; Hassan, A.; Abdel-Wahab, M.F.; Wissa, J. Urinary Schistosomiasis Treated with Sodium Antimony Tartrate--a Quantitative Evaluation. BMJ 1968, 3, 713–714. https://doi.org/10.1136/bmj.3.5620.713. 348. Gopalratnam, P.C.; Mason, N.S.; Sparks, R.E. Microencapsulation of astiban acid for the treatment ofSchistosomiasis mansoni. Appl. Biochem. Biotechnol. 1984, 10, 213–220. https://doi.org/10.1007/bf02783753. (In English) 349. de Melo, A.L.; Silva-Barcellos, N.M.; Demicheli, C.; Frézard, F. Enhanced schistosomicidal efficacy of tartar emetic encapsulated in pegylated liposomes. Int. J. Pharm. 2003, 255, 227–230. https://doi.org/10.1016/s0378-5173(03)00125-x. (In English) 350. Meyerhoff, A. U.S. Food and Drug Administration Approval of AmBisome (Liposomal Amphotericin B) for Treatment of Visceral Leishmaniasis. Clin. Infect. Dis. 1999, 28, 42–48; discussion 49–51. https://doi.org/10.1086/515085. (In English) 351. Croft, S.L.; Coombs, G.H. Leishmaniasis—Current chemotherapy and recent advances in the search for novel drugs. Trends Parasitol. 2003, 19, 502–508. https://doi.org/10.1016/j.pt.2003.09.008. (In English) 352. Alvar, J.; Cañavate, C.; Gutiérrez-Solar, B.; Jiménez, M.; Laguna, F.; López-Vélez, R.; Molina, R.; Moreno, J. Leishmania and human immunodeficiency virus coinfection: The first 10 years. Clin. Microbiol. Rev. 1997, 10, 298–319. https://doi.org/10.1128/cmr.10.2.298. (In English) 353. Rosenthal, E.; Marty, P.; Poizot-Martin, I.; Reynes, J.; Pratlong, F.; Lafeuillade, A.; Jaubert, D.; Boulat, O.; Dereure, J.; Gambarelli, F.; et al. Visceral leishmaniasis and HIV-1 co-infection in southern France. Trans. R. Soc. Trop. Med. Hyg. 1995, 89, 159–162. https://doi.org/10.1016/0035-9203(95)90476-x. (In English) 354. Bryceson, A.; Chulay, J.; Ho, M.; Mugambii, M.; Were, J.; Muigai, R.; Chunge, C.; Gachihi, G.; Meme, J.; Anabwani, G.; et al. Visceral leishmaniasis unresponsive to antimonial drugs I. Clinical and immunological studies. Trans. R. Soc. Trop. Med. Hyg. 1985, 79, 700–704. https://doi.org/10.1016/0035-9203(85)90197-x. (In English) 355. Davidson, R.N.; Russo, R. Relapse of Visceral Leishmaniasis in Patients Who Were Coinfected with Human Immunodeficiency Virus and Who Received Treatment with Liposomal Amphotericin B. Clin. Infect. Dis. 1994, 19, 560. https://doi.org/10.1093/clinids/19.3.560. (In English) 356. Lopez-Velez, R.; Perez-Molina, J.A.; Bellas, C.; Perez-Corral, F.; Villarrubia, J.; Guerrero, A.; Escribano, L.; Baquero, F.; Alvar, J. Clinicoepidemiologic characteristics, prognostic factors, and survival analysis of patients coinfected with human immunodeficiency virus and Leishmania in an area of Madrid, Spain. Am. J. Trop. Med. Hyg. 1998, 58, 436–443. https://doi.org/10.4269/ajtmh.1998.58.436. (In English) 357. Vianna, G. Tratamento da leishmaniose tegumentar por injecoes intravenosas de tartaro emetic. Arq. Bras. Med. 1912, 4, 426–428. 358. Di Cristina, G.; Caronia, G. Sulla terapia della leishmaniosi interna. Pathologica 1915, 7, 82–83. 359. Cook, G. Leonard Rogers KCSI FRCP FRS (1868–1962) and the founding of the Calcutta School of Tropical Medicine. Notes Rec. R. Soc. J. Hist. Sci. 2006, 60, 171–181. https://doi.org/10.1098/rsnr.2006.0146. 360. Brahmachari, U.N. Chemotherapy of antimonial compounds in kala-azar infection. Part IV. Further observations on the therapeutic values of urea stibamine. By U.N. Brahmachari, 1922. Indian J. Med. Res. 1989, 89, 393–404. (In English) Int. J. Environ. Res. Public Health 2022, 19, 4669 29 of 30 361. Shortt, H.E. Recent research on kala-azar in India. Trans. R. Soc. Trop. Med. Hyg. 1945, 39, 13–31. https://doi.org/10.1016/00359203(45)90030-7. 362. Berman, J.D. Chemotherapy for Leishmaniasis: Biochemical Mechanisms, Clinical Efficacy, and Future Strategies. Clin. Infect. Dis. 1988, 10, 560–586. https://doi.org/10.1093/clinids/10.3.560. (In English) 363. Holmes, F. Mass Treatment of Oriental Sores. J. R. Army Med. Corps 1937, 69, 258–260. https://doi.org/10.1136/jramc-69-04-06. 364. Andrews, L.A. Three Cases of Tropical Sore. J. R. Army Med. Corps 1923, 40, 371–372. https://doi.org/10.1136/jramc-40-05-09. 365. Roberts, W.L.; Berman, J.D.; Rainey, P.M. In vitro antileishmanial properties of tri- and pentavalent antimonial preparations. Antimicrob. Agents Chemother. 1995, 39, 1234–1239. https://doi.org/10.1128/aac.39.6.1234. (In English) 366. Sereno, D.; Lemesre, J.L. Axenically cultured amastigote forms as an in vitro model for investigation of antileishmanial agents. Antimicrob. Agents Chemother. 1997, 41, 972–976. https://doi.org/10.1128/aac.41.5.972. (In English) 367. Sereno, D.; Cavaleyra, M.; Zemzoumi, K.; Maquaire, S.; Ouaissi, A.; Lemesre, J.L. Axenically Grown Amastigotes of Leishmania infantum Used as an In Vitro Model To Investigate the Pentavalent Antimony Mode of Action. Antimicrob. Agents Chemother. 1998, 42, 3097–3102. https://doi.org/10.1128/aac.42.12.3097. (In English) 368. Roberts, W.L.; Rainey, P.M. Antileishmanial activity of sodium stibogluconate fractions. Antimicrob. Agents Chemother. 1993, 37, 1842–1846. https://doi.org/10.1128/aac.37.9.1842. (In English) 369. Callahan, H.L.; Portal, A.C.; Devereaux, R.; Grogl, M. An axenic amastigote system for drug screening. Antimicrob. Agents Chemother. 1997, 41, 818–822. https://doi.org/10.1128/aac.41.4.818. (In English) 370. Ephros, M.; Bitnun, A.; Shaked, P.; Waldman, E.; Zilberstein, D. Stage-Specific Activity of Pentavalent Antimony against Leishmania donovani Axenic Amastigotes. Antimicrob. Agents Chemother. 1999, 43, 278–282. https://doi.org/10.1128/aac.43.2.278. (In English) 371. Service, M.W. Tsetse flies (Order Diptera: Family Glossinidae). In A Guide to Medical Entomology; Macmillan Education: London, UK, 1980; pp. 95–101. 372. Vanhamme, L.; Pays, E. The trypanosome lytic factor of human serum and the molecular basis of sleeping sickness. Int. J. Parasitol. 2004, 34, 887–898. https://doi.org/10.1016/j.ijpara.2004.04.008. (In English) 373. Bronner, U.; Doua, F.; Ericsson, Örjan; Gustafsson, L.L.; Miézan, T.; Rais, M.; Rombo, L. Pentamidine concentrations in plasma, whole blood and cerebrospinal fluid during treatment of Trypanosoma gambiense infection in Côte d’Ivoire. Trans. R. Soc. Trop. Med. Hyg. 1991, 85, 608–611. https://doi.org/10.1016/0035-9203(91)90364-5. (In English) 374. Wenzler, T.; Yang, S.; Braissant, O.; Boykin, D.W.; Brun, R.; Wang, M.Z. Pharmacokinetics, Trypanosoma brucei gambiense Efficacy, and Time of Drug Action of DB829, a Preclinical Candidate for Treatment of Second-Stage Human African Trypanosomiasis. Antimicrob. Agents Chemother. 2013, 57, 5330–5343. https://doi.org/10.1128/aac.00398-13. (In English) 375. Sharma, S.; Anand, N. Chapter 4—Organometaliics. In Pharmacochemistry Library; Sharma, S.; Anand, N.; Eds.; Elsevier: Amsterdam, The Netherlands, 1997; Volume 25, pp. 124–147. 376. Thétiot-Laurent, S.A.-L.; Boissier, J.; Robert, A.; Meunier, B. Schistosomiasis Chemotherapy. Angew. Chem. Int. Ed. 2013, 52, 7936–7956. https://doi.org/10.1002/anie.201208390. 377. Chitsulo, L.; Engels, D.; Montresor, A.; Savioli, L. The global status of schistosomiasis and its control. Acta Trop. 2000, 77, 41–51. https://doi.org/10.1016/s0001-706x(00)00122-4. (In English) 378. Fenwick, A.; Rollinson, D.; Southgate, V. Implementation of Human Schistosomiasis Control: Challenges and Prospects. Adv. Parasitol. 2006, 61, 567–622. https://doi.org/10.1016/s0065-308x(05)61013-5. (In English) 379. Gryseels, B.; Polman, K.; Clerinx, J.; Kestens, L. Human schistosomiasis. Lancet 2006, 368, 1106–1118. https://doi.org/10.1016/s0140-6736(06)69440-3. (In English) 380. Manson-Bahr, P. Manson’s Tropical Diseases. A Manual of the Diseases of Warm Climates, 14th ed.; Williams & Wilkins: Baltimore, MD, USA, 1954; p. 1144. 381. Goodman, L.S.; Gilman, A. The Pharmacological Basis of Therapeutics, 2nd ed.; Macmillan: New York, NY, USA, 1956. 382. Newham, Trypanosomiasis in the East African Campaign. J. R. Army Med. Corps 1919, 33, 299–311. https://doi.org/10.1136/jramc33-04-02. 383. Jopling, W.H. The eradication of schistosomiasis; a plea for a rational approach to the problem. J. Trop. Med. Hyg. 1949, 52, 121– 126. (In English) 384. Mainzer, F.; Krause, M. Changes of the electrocardiogram appearing during antimony treatment. Trans. R. Soc. Trop. Med. Hyg. 1940, 33, 405–418. https://doi.org/10.1016/s0035-9203(40)90042-6. 385. Mishra, M.; Biswas, U.; Jha, A.; Khan, A. Amphotericin versus sodium stibogluconate in first-line treatment of Indian kala-azar. Lancet 1994, 344, 1599–1600. https://doi.org/10.1016/s0140-6736(94)90406-5. (In English) 386. Thakur, C.P.; Sinha, G.P.; Pandey, A.K. Comparison of regimens of amphotericin B deoxycholate in kala-azar. Indian J. Med. Res. 1996, 103, 259–263. (In English) 387. Bryceson, A. A policy for leishmaniasis with respect to the prevention and control of drug resistance. Trop. Med. Int. Health 2001, 6, 928–934. https://doi.org/10.1046/j.1365-3156.2001.00795.x. 388. Peraza, M.A.; Ayala-Fierro, F.; Barber, D.S.; Casarez, E.; Rael, L.T. Effects of micronutrients on metal toxicity. Environ. Health Perspect. 1998, 106, 203–216. https://doi.org/10.1289/ehp.98106s1203. (In English) 389. Kabata-Pendias, A.; Mukherjee, A.B. Trace Elements from Soil to Human; Springer: New York, NY, USA, 2007. 390. Christopoulou, A.; Dimitriou, E. Impacts of climate chance scenarios on Spercheios River hydrology. Presented at the 11th International Hydrogeological Congress of Greece, Athens, Greece, 4–6 October 2017. Int. J. Environ. Res. Public Health 2022, 19, 4669 30 of 30 391. Stefanidis, K.; Christopoulou, A.; Poulos, S.; Dassenakis, E.; Dimitriou, E. Nitrogen and Phosphorus Loads in Greek Rivers: Implications for Management in Compliance with the Water Framework Directive. Water 2020, 12, 1531. Available online: https://www.mdpi.com/2073-4441/12/6/1531 (accessed on 13 January 2022). 392. Matschullat, J.; Ottenstein, R.; Reimann, C. Geochemical background—Can we calculate it? Environ. Geol. 2000, 39, 990–1000. https://doi.org/10.1007/s002549900084. 393. Reimann, C.; Garrett, R.G. Geochemical background—Concept and reality. Sci. Total Environ. 2005, 350, 12–27. https://doi.org/10.1016/j.scitotenv.2005.01.047. 394. Nastos, P.T.; Paliatsos, A.G.; Anthracopoulos, M.B.; Roma, E.S.; Priftis, K.N. Outdoor particulate matter and childhood asthma admissions in Athens, Greece: A time-series study. Environ. Health 2010, 9, 45. https://doi.org/10.1186/1476-069x-9-45. 395. Samoli, E.; Nastos, P.T.; Paliatsos, A.G.; Katsouyanni, K.; Priftis, K.N. Acute effects of air pollution on pediatric asthma exacerbation: Evidence of association and effect modification. Environ. Res. 2011, 111, 418–424. https://doi.org/10.1016/j.envres.2011.01.014. 396. Kelepertzis, E.; Argyraki, A.; Botsou, F.; Aidona, E.; Szabó, Á.; Szabó, C. Tracking the occurrence of anthropogenic magnetic particles and potentially toxic elements (PTEs) in house dust using magnetic and geochemical analyses. Environ. Pollut. 2019, 245, 909–920. https://doi.org/10.1016/j.envpol.2018.11.072.