Next Article in Journal
CISD1 Is a Breast Cancer Prognostic Biomarker Associated with Diabetes Mellitus
Next Article in Special Issue
Non-Uniform Bioaccumulation of Lead and Arsenic in Two Remote Regions of the Human Heart’s Left Ventricle: A Post-Mortem Study
Previous Article in Journal
Short-Term Consequences of Single Social Defeat on Accumbal Dopamine and Behaviors in Rats
Previous Article in Special Issue
Joint Action Toxicity of Arsenic (As) and Lead (Pb) Mixtures in Developing Zebrafish
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Biological Effects of Human Exposure to Environmental Cadmium

by
Massimiliano Peana
1,*,
Alessio Pelucelli
1,*,
Christos T. Chasapis
2,
Spyros P. Perlepes
3,
Vlasoula Bekiari
4,
Serenella Medici
1 and
Maria Antonietta Zoroddu
1
1
Department of Chemical, Physical, Mathematical and Natural Sciences, University of Sassari, 07100 Sassari, Italy
2
Institute of Chemical Biology, National Hellenic Research Foundation, 11635 Athens, Greece
3
Department of Chemistry, University of Patras, 26500 Patras, Greece
4
School of Agricultural Science, University of Patras, 30200 Messolonghi, Greece
*
Authors to whom correspondence should be addressed.
Biomolecules 2023, 13(1), 36; https://doi.org/10.3390/biom13010036
Submission received: 24 October 2022 / Revised: 20 December 2022 / Accepted: 21 December 2022 / Published: 24 December 2022
(This article belongs to the Special Issue Toxic and Essential Metals in Human Health and Disease 2022-2023)

Abstract

:
Cadmium (Cd) is a toxic metal for the human organism and for all ecosystems. Cd is naturally found at low levels; however, higher amounts of Cd in the environment result from human activities as it spreads into the air and water in the form of micropollutants as a consequence of industrial processes, pollution, waste incineration, and electronic waste recycling. The human body has a limited ability to respond to Cd exposure since the metal does not undergo metabolic degradation into less toxic species and is only poorly excreted. The extremely long biological half-life of Cd essentially makes it a cumulative toxin; chronic exposure causes harmful effects from the metal stored in the organs. The present paper considers exposure and potential health concerns due to environmental cadmium. Exposure to Cd compounds is primarily associated with an elevated risk of lung, kidney, prostate, and pancreatic cancer. Cd has also been linked to cancers of the breast, urinary system, and bladder. The multiple mechanisms of Cd-induced carcinogenesis include oxidative stress with the inhibition of antioxidant enzymes, the promotion of lipid peroxidation, and interference with DNA repair systems. Cd2+ can also replace essential metal ions, including redox-active ones. A total of 12 cancer types associated with specific genes coding for the Cd-metalloproteome were identified in this work. In addition, we summarize the proper treatments of Cd poisoning, based on the use of selected Cd detoxifying agents and chelators, and the potential for preventive approaches to counteract its chronic exposure.

1. Introduction

Cadmium (Cd) was discovered in 1817 in Germany by Friedrich Stromeyer as an impurity in zinc carbonate (ZnCO3). The name derives from the Latin word cadmia and the Greek word καδμεία, an older name for the common zinc (Zn) ore, calamine. Cd is naturally present in the Earth’s crust, at an average concentration estimated to be between 0.1 and 0.2 parts per million (ppm) [1]. Atmospheric Cd can form as a result of natural activities, such as spontaneous biomass combustion and volcanic eruptions. Despite its natural presence on the planet, no biological function has been found for Cd in higher organisms [2], while its toxicity is well known and has been the subject of numerous studies [3]. Cd continues to occur in the environment, particularly in aerosol form, as a result of human activities, such as the burning of fossil fuels and wastes, or the process of mining metal ores and industrial emissions, representing a current threat to public health [4]. Humans are protected from chronic exposure to low Cd concentrations by the presence of metallothioneins (MTs), a family of ubiquitous small cysteine-rich proteins, the specific function of which is to regulate the metabolism of Zn. MTs play important roles in protection against ion toxicity from several heavy metals, DNA damage, and oxidative stress. Thanks to the presence of many sulfhydryl groups (–SH), MTs are able of complexing, under tolerable exposures, almost all ingested Cd ions [2]. In the kidneys, the resulting Cd2+-complexes are, in part, excreted with the urine. The cysteine content of MTs can be up to 30%. These residues can be arranged in motifs (e.g., Cys-x-Cys or Cys-x-x-Cys), and this aspect is fundamental for the formation of metal-binding clusters. The protective effect of MTs against metal toxicity has been largely discussed in a number of studies [5,6]. Apart from Cd, MTs can bind other metal ions, such as mercury, platinum, etc., in order to protect cells and tissues against their toxicity [7]. Mutant flies with increased quantities of MTs in their genes are able to survive on a diet rich in heavy metals, while normal flies would die from the same diet [8]. An increased expression of these proteins can be preventive against heavy metal intoxication: this expression is regulated by the metal regulatory transcription factor 1 (MTF-1).
If the amount of Cd absorbed by the human body exceeds the complexation capacity of MTs, then there will be an accumulation of the metal, mainly in the kidneys (30%) and liver (30%), with the rest dispersed throughout the other organs and having an extremely long half-life of 10–30 years [9,10]. The biological half-life of Cd in the blood has been estimated to be in the range of 75 to 128 days; however, this half-life mainly represents its deposition in the organs, not the body’s clearance rate [10,11]. Cd bioaccumulation in various tissues affects cell functionality, such as proliferation and differentiation, and is responsible for oxidative stress by the generation of reactive oxygen species (ROS), despite its inability to generate free radicals directly, instead forming them via the replacement of redox-active metal ions, such as Fe2+ and Cu2+, from their metal-binding sites in proteins. The resulting oxidative stress will induce cellular damage and apoptosis [12]. Cd binding into the mitochondria can inhibit both cellular respiration and oxidative phosphorylation, even at low concentrations [13,14]. With these oxidative mechanisms, Cd has been shown to interfere with almost all major DNA repair systems. It is indeed able to impair nucleotide excision repair (NER), base excision repair (BER), and mismatch repair (MMR) [15]. The disturbance of DNA repair processes may explain the co-mutagenic effects in combination with other carcinogenic sources, as DNA repair systems are not only required for the repair of DNA-induced damage but also for the removal of DNA lesions due to endogenous processes and to keep replication errors low [16]. Once absorbed, the thiol groups of cysteines appear to be the critical targets of Cd ions in proteins, enzymes, and endogenous antioxidants such as glutathione (GSH, L-γ-glutamyl-L-cysteinyl-glycine). The inhibition of antioxidant defense enzymes, such as superoxide dismutase (SOD), lactate dehydrogenase (LDH), catalase (CAT), thioredoxin reductase (TrxR), and glutathione peroxidase (GPx) leads to the further dysregulation of the cellular redox state [3,12,17]. The present paper considers exposure and potential health concerns on environmental cadmium, with particular attention to its carcinogenic action. Moreover, we summarize the strategy of environmental Cd remediation and the proper treatments of Cd poisoning, based on the use of selected detoxifying agents and chelators, either alone or in combination.

2. The Bioinorganic Chemistry of Cadmium

Cd (electronic configuration (Kr) 4d10 5s2) belongs to group 12 of the periodic table, together with zinc and mercury, but it is chemically more similar to the former than to the latter. It can be found with oxidation states 0, +1, and +2; however, only Cd2+ ions are stable under normal conditions, as well as for Zn. Unlike mercury (Hg), which, with simple anions, form compounds having a covalent character, those of Cd have mainly ionic characteristics. Cd2+ has an ionic radius of 0.97 Å, very similar to those of Ca2+ (0.99 Å) and Na+ (1.09 Å), facilitating their replacement in biological matrices such as bones. Cd2+ can form coordination complexes with ligands and, particularly, with biological ones such as proteins and nucleic acids. These interactions are modulated by the innate chemical features of Cd2+, which is a soft metal ion in the hard/soft acid-base (HSAB) classification [18,19]. The stereochemistry of Cd2+ complexes with ligands in biological fluids is useful for understanding the biological behavior and fate of Cd2+ ions. According to the HSAB model, Cd2+ is a soft acid having a preference for sulfhydryl donors (soft bases), whereas hard carboxylate/carbonyl aminoacidic groups will generally prefer the borderline Zn2+, despite the chemical similarities between these ions [19]. The competition between Cd2+ and other divalent cations (Zn2+, Mn2+, Fe2+, Ni2+, and Cu2+) during complex formation with amino acids, peptides, and chelating agents has been reviewed previously [20].
Cd2+ can have a wide range of coordination numbers (from 2 to 8), but it mostly shows the coordination numbers 4 or 6, represented, respectively, by regular tetrahedral and octahedral structures with proteins (Figure 1a,b). Depending on the protein to which Cd is bound, clusters can also be generated. MTs are a lightning example of such polynuclear complexes, as cysteine-rich regions of the proteins are able to cluster Cd2+ (Figure 1c).
The geometry distribution of Cd2+ in the X-ray structures of those proteins deposited in the Research Collaboratory for Structural Bioinformatics Protein Data Bank (RCSB PDB https://www.rcsb.org/ accessed on 22 October 2022) in which Cd2+ forms mononuclear complexes is shown in Figure 2 and relative data are reported in Table S1 of the Supplementary Materials. From the analysis of the solved crystal structure available in the database, it transpires that the main coordination geometries around Cd2+ are represented by octahedral (43.0%), trigonal bipyramidal (26.7%), and tetrahedral (15.9%) forms (derived geometries were included in the count) [24].
Part of the toxicity of Cd2+ is due to its ability to mimic other divalent ions, particularly Zn2+ and Ca2+, which have a similar radius and a similar valence.
A useful tool to analyze the possibilities that a metal ion has to replace another one in its natural metal coordination sites, is based on the use of a representative minimal functional site (MFS) [25]. The MFS describes the local 3D environment around the metal ion and is independent of the larger context of the protein fold in which it is embedded. In summary, MFSs do not depend on the overall macromolecular structure but instead represent a region crucial for the correct physiological function of a selected metalloprotein [25]. It is possible to extract statistical data from the analysis of MFSs (through the 3D X-ray structures of metal–proteins deposited in the Protein Data Bank), by counting how many amino acidic patterns are common to the different ligands in their coordination binding sites. It is then possible, by crossing these data with the protein selectivity for a given metal ion, to understand how probable the substitution is of the original metal ion with another metal ion. In the case of Cd2+, on the basis of data from Metalpdb, the database of metal sites in biological macromolecular structures (https://metalpdb.cerm.unifi.it/, accessed on 22 October 2022) [24], the analysis of the possible cadmium substitution for other metal sites yields the results shown in Figure 3 and Table S2 of the Supplementary Materials.
Interestingly, the statistical analysis based on a total of 150,149 PDB structures showed that Cd2+ has the probability of substituting for other metal ions in 4189 proteins (~2.8% of the total number of PDB structures). In particular, Cd2+ has been shown to have a major affinity for the Zn2+ binding sites (1245 proteins, 29.7%), Ca2+ (590 proteins, 14.08%), and then Mg2+ (610 proteins, 14.9%). While it is widely reported that Cd2+ can substitute Zn2+ at its metal coordination sites as well as Ca2+, the replacement of Mg2+ by Cd2+ is not well known, nor has it been well studied, due to the huge variety of roles that Mg plays in the human body. It has been suggested that cadmium can interfere with the absorption of magnesium in the gastrointestinal tract, affecting its homeostasis. Conversely, several reports suggest that enhanced dietary magnesium intake can mitigate the pathogenic impact of cadmium exposure and its induced alterations in the homeostasis of zinc, copper, and magnesium itself [26,27].

3. Uses and Environmental Dispersion of Cadmium

After its first isolation, Cd was used for a variety of processes, firstly as a painting and plating compound (cadmium sulfide, CdS, was used as a yellowish pigment, and cadmium selenide, CdSe, was used as a red one) for about 100 years, mainly because of the particular pigment’s brightness. Cadmium pigments have been used by artists since the 19th century. Because of their resistance to temperatures up to 3000 °C, these Cd-based pigments can be used for coloring hot pipes or glass; examples include red traffic lights and the lit-up stars on the Moscow Kremlin. During the last years CdS and CdSe semiconductor nanomaterials due to their photoluminescent properties have been extensively studied for their use in various environmental monitoring applications, photovoltaic cells as sensitizers, as well as bio-imaging and nanomedicine [28,29]. The majority of Cd has been used (and still is used) in battery technology, in particular, rechargeable nickel-cadmium (Ni–Cd) batteries. The invention of the rechargeable Ni–Cd battery goes back to 1899 and has played a significant role in electrical technology during the twentieth century. In 2002, the European Union put a limit on the use of Cd in electronics, repealing this action in 2016 and establishing a maximum content of Cd of 0.1%, reducing the limit on Cd MCV (maximum concentration value) in homogeneous material to 0.01% [30]. Unfortunately, only a few countries have restrictions regarding the exportation of Cd components, and even fewer have laws about its recycling. In developing countries, this is becoming a huge problem for public health, where electronic waste (e-waste) recycling activities are often conducted without all the safety measures that are needed [31,32]. Cd was a widely used component of plastics, especially PVC, where it works as a stabilizer, but now it has been completely replaced with less harmful metals and alloys, such as barium–zinc alloy. Another source of cadmium for half a century was from its use as a metrology standard from 1907 to 1960. During that period, the angstrom was defined by fixing the wavelength of the distinct, red spectral line of cadmium at 6438.4696 Å. Due to its ability to efficiently capture neutrons, the metal played a remarkable role in the development of the first nuclear reactors; cadmium-coated rods were used to control the nuclear reaction.
Large amounts of Cd are obtained as byproducts of zinc smelting, as the two metals are often naturally associated in minerals. Therefore, many of the environmental contaminations from Cd occur in areas close to the treatment and smelting plants of zinc, but also those of copper and lead. Cd is released into the environment during the disposal and incineration phases of waste containing, in particular, plastic materials, steel plated with cadmium, and nickel-cadmium batteries. Cd exposure in the workplace takes place during mining and work using Cd-containing ores [33].

4. Human Exposure to Cadmium

4.1. Ingestion

Cd tends to persist and accumulate in the soil and then enters the metabolism of plants. Its accumulation in edible plant parts, including fruits and seeds, leads to Cd’s entry into the food chain [34]. This accumulation increases with the decrease in pH in soil; consequently, acid rain has the effect of increasing Cd concentrations in plants. The foods that mainly contribute to the daily Cd intake in Western countries include cereals and bread (34%), leafy vegetables, in particular spinach, among adults (20%), potatoes (11%), legumes and nuts (7%), stem/root vegetables (6%), and fruits (5%). In terms of Eastern countries, fish and shellfish can be identified as the major Cd sources, in addition to grains and vegetables, which are represented particularly by rice [35]. A particular danger may be represented by rice grains, which account for more than half of the total Cd intake in Eastern regions of the world, at 44% in Japan [36] and 56% in China [37]. The typical dietary Cd intake has been estimated to be about 30–50 µg/day [38], but normal individuals absorb only a small portion of an orally ingested dose (1–10%) [39]. Although the health risk from dietary cadmium exposure in Eastern nations is generally low, it still remains a cause for concern for some subgroups. In fact, being disseminated across the planet, there are areas with very high levels of Cd in the soil. The crop uptake of Cd in these areas can lead to significant dietary exposures for the people living nearby. For example, in the Jinzu and Kakehashi river basins in Japan, there are areas with soil that is heavily contaminated by Cd, derived from industrial waste [40,41]. Local people who frequently consumed rice cooked with Cd-contaminated water, developed a severe kidney and bone syndrome called “Itai-Itai” disease, characterized by bone deformation and multiple fractures, especially in women [42]. The radiographs showed the presence of osteomalacia and bone decalcification, as well as osteoporosis, and a series of bone deformities due to the replacement of Ca2+ by Cd2+, with a consequent alteration of the normal bone structure.
It is important to note that in the US, in the late 1960s, the average Cd consumption was estimated to be around 26 μg/day/person [43], and, in the first years of the 1990s, was calculated to be around 18.9 μg/day/person [44]; now, the human Cd intake has been lowered to 4.63 μg/day/person, on average. This corresponds to a value of 75% under the tolerable limit of toxicity. The diminishing of Cd intake in the last 50 years could be attributed to the lowered activities of leaking sewage sludge into agricultural soil, because of more efficient control and environmental awareness in developed countries. These activities were mainly responsible for the transfer of Cd, adsorbed by plants, into the food chain, contributing to increasing human exposure to the metal. Some aquatic organisms can also be largely affected by Cd accumulation reaching levels above regulatory standards. While muscle tissue in fish does not represent a site of accumulation of Cd [45], in commercial oysters (Crassostrea gigas) from southern Korea, the mean Cd concentration of 80 samples was found to be 0.591 μg/kg, which was much higher than the mean Cd concentration in water samples, 0.0021 μg/L [46]. The acidification of seawater could increase the accumulation of Cd by most common seafood species, such as Mytilus edulis, Tegillarca granosa, and Meretrix meretrix, with a potential threat to seafood safety [47,48]. Several protein transporters have been identified as having a role in cadmium uptake in the intestinal tract. These include divalent metal transporter 1 (DMT1), cationic amino acid transporter 1 (CaT1), zinc transporters (ZIP4, ZIP8, ZIP14, and ZnT1), copper-transporting P-type ATPase or Menkes ATPase (ATP7A), calcium channel TRVP6, and metallothionein MT-1 and MT-2, for which the increased Cd oral concentrations resulted in relatively increased gene expression [5]. Once introduced into the blood, Cd reaches various organs, such as the kidneys and skeleton, through systemic distribution. Initially, Cd in plasma is bound with high molecular-mass proteins such as albumin, then it is found bound with proteins of the molecular size of MTs, which are believed to be responsible for transporting Cd to the kidneys [5].

4.2. Inhalation

Cd air levels can be hundreds of times greater in the workplace than in the general environment. For example, the Occupational Safety and Health Administration (OSHA) fixed the permissible exposure limit (PEL) of Cd fume or Cd oxide in the workplace at 0.1 mg/m3, whereas concentrations of Cd in ambient air are 1 × 10−6 mg/m3 in non-industrialized areas and 4 × 10−5 mg/m3 in urban areas, respectively. Non-occupational Cd exposure from the air is not expected to pose the risk of adverse health effects. In general, Cd air levels are usually not sufficient to cause health problems among the general population. Even in those areas with high industrial emissions of Cd, its average atmospheric concentration is not higher than 35 ng Cd/m3 of air. A Brazilian study showed higher blood Cd levels (~0.22 μg/L) in automotive battery-manufacturing workers than in the control group, who had a mean blood Cd level of around 0.03 μg/L. However, the mean blood Cd level in battery workers was even lower compared to the WHO standard (10 μg/L) [49]. In other words, elevated occupational exposure to Cd decreased in the last 50 years, and such a change is due to better regulations regarding Cd exposure in the workplace. Therefore, the daily deposition rate would be 0.175 ng Cd, assuming that 20 m3 of air would be inhaled per day, with a deposition rate of 25% [50]. However, Cd exposure is still considered a threat in workplaces in developing countries, where safety standards are not regulated or respected. The major routes of Cd exposure occurring in the general population are from the ingestion of Cd-contaminated foods and cigarette smoking, since the tobacco plants take up Cd from polluted soil, due to its similarity to Zn [51]. Compared to non-smokers, Cd levels are four to five times higher in the blood and two to three times higher in the kidneys for tobacco smokers [52,53]. Tobacco plants accumulate every polluting metal that might be present in the neighborhood, and, in particular, arsenic (As) since some of its compounds are widely used in the production of pesticides or herbicides [54]. Since both As and Cd are classified by the IARC as belonging to Group 1 of the carcinogenic substances, Cd is believed to be linked to lung cancer in smokers, in synergism with As, as well as with other potentially toxic substances that are eventually present in tobacco leaves.

4.3. Permeation

There are negligible amounts of Cd absorbed through the skin; thus, it is not considered to be a critical route of exposure. However, recent research highlighted the environmental significance of photosensitive CdS and CdSe pigments and nano semiconductors, whose oxidized products (cadmium sulfate, CdSO4, and cadmium selenite, CdSeO4) are considerably more soluble and bioavailable and are, consequently, potentially more dangerous [55]. An in vitro study using human full-thickness skin as a model to characterize the impact of Cd exposure on skin showed that the metal penetrates only the epidermis; it was shown before that its solubility into the stratum corneum layer is a rate-limiting process [56]. Permeability is affected by the Cd concentration applied to the skin. It may also be influenced by pH and metal speciation [19]. The high concentrations of Cd in the epidermis may explain the induction of MTs, as Cd is very effective in activating their expression [56].

5. Effect of Human Exposure to Cadmium

5.1. General Effects That Are Harmful to Health

In this section, we will describe the main health adverse effects of cadmium exposure, while in the next section, the strictly carcinogenic aspects will be explored. Figure 4 illustrates the main outcomes in health effects following chronic cadmium exposure.
The organs most affected by Cd toxicity are the kidneys; as much as about 30% of body Cd is deposited in the kidney tubule region, provoking tubular damage proportionally to the quantity of Cd not bound to MTs. In one study, diabetics were more susceptible to renal tubular damage from Cd exposure than were the controls [57]. Another common disease that can be correlated to Cd exposure is osteoporosis and/or osteomalacia. Cd has a deleterious impact on bones due to its impairment of Vitamin D uptake in the kidneys. Cd also prevents the absorption of calcium at the gut level, causing general bone disease, such as the aforementioned Itai-Itai disease [58]. Cd has a remarkable role in the reproductive system, where the metal interferes in a number of ways [59]. It has been shown that Cd can alter cell adhesion in the testis, interfering with the normal migration of germ cells across the seminiferous epithelium. Furthermore, decreased testicular growth rate, plasma testosterone, and reduced sperm count and motility have been linked to Cd-induced oxidative stress, as the result of a decrease in GPx, CAT, mitochondrial Mn–SOD, and cytosolic CuZn–SOD [60].
Another organism district that is affected by Cd toxicity is the cardiovascular system. Cd can cause hypertension by the inhibition of the endothelial nitric oxide synthase and the suppression of acetylcholine-induced vascular relaxation [61,62]. Cd could also affect glucose metabolism by acting on a variety of different organs, such as the pancreas, liver, and adrenal gland. Studies suggest a direct effect of Cd on the pancreas, and there is evidence that Cd can alter insulin release from pancreatic β-cells while increasing the activity of all the four enzymes responsible for gluconeogenesis [63]. As has been noticed in Itai-Itai disease, Cd also seems to induce anemia, due to the suppression of erythropoietin production [64]; this mechanism, linked to the suppression of iron transport in the duodenum, may cause iron-deficient anemia [65]. Lungs are also a target organ for Cd toxicity. In chronic Cd exposure, progressive pulmonary fibrosis and impaired lung function with obstructive lung disease may occur [66,67]. Cd also induces neurological dysfunction and brain toxicity, with a complex mechanism [68,69].
Cd could be considered a weak genotoxic and mutagenic agent, due to its low affinity with DNA [70]. A number of toxicogenomic studies confirm the involvement of Cd in the mutation of the following genes: immediate early response genes (IEGs), stress response genes, transcriptional factors, and translational factors [12]. One of the mechanisms by which Cd influences gene expression involves not only the homeostasis of Ca, an element that can be directly mimicked by Cd, but also an element whose concentration inside the cell can be influenced by it [71].

5.2. Carcinogenicity of Cadmium

One of the major Cd-induced mechanisms of carcinogenesis is oxidative stress [12]; this is, in part, caused, as mentioned above, by changes in Ca concentrations inside the cell. Cd cannot directly produce ROS because it does not undergo Fenton-like reactions, but, on the other hand, it can replace redox-active metal ions and can inhibit the activity of antioxidant enzymes, as well as promote lipid peroxidation [72]. The major enzymatic antioxidant is SOD, which degrades O2, and the CAT and GSH redox system, which inactivates H2O2 and hydroperoxides. Three forms of SOD may be important: Mn-SOD (which is located in the mitochondria), Cu–Zn SOD (which resides in the cytoplasm), and extracellular SOD (which lines the blood vessels). GSH, present in high concentrations in every cell, is able to detoxify Cd in the human blood cell, and its synthesis (triggered by the transcription factor Nrf2) is enhanced after Cd exposure [73]. Cd has been shown to impair global genome nucleotide excision repair (GG-NER): it can interfere with the removal of DNA lesions in cultured mammalian cells caused by benzo[a]pyrene and UVC [74]. The underlying mechanism in this impairment has been identified as an interaction with the Zn-binding proteins, which show a common motif (Zn finger), where Zn is complexed to four cysteine and/or histidine residues [16,75,76]. Cd2+ can substitute Zn2+ in these metal binding sites, inactivating the proteins. Another process of the DNA repair pathway is base excision repair (BER), but it is different from NER. BER, in fact, is activated by a specific class of DNA repair enzymes called glycosylases. Low concentrations of Cd disturb the repair of oxidative DNA base damage, as well as DNA alkylation damage, in mammalian cells [77]. It is important to cite another mechanism that is relevant for maintaining genomic stability; the mismatch repair (MMR) is responsible for the repair of mismatched bases after DNA replication. Cells deficient in MMR usually tend to mutate often and are associated with a greater risk of developing several types of cancer. MMR also has a role in apoptosis, and it has been shown that MMR-deficient cells are about 100 times more resistant to the cytotoxicity of alkylating agents [78]. In extracts of human cells, Cd inhibited at least one step of the MMR process [79]. The underlying mechanism in this inhibition seems to be the interaction of Cd in the process of ATP binding the MMR enzymes, plus their hydrolysis, reducing their binding activity to DNA bases and interfering with their ability to recognize mismatched DNA base pairings [80,81].
Cd intoxication can also lead to cell death by apoptosis with different pathways [82,83]. The extrinsic pathway is initiated by binding the cytokine ligands, such as the Fas ligand (FasL) and the tumor necrosis factor alpha (TNF-ɑ), along with the death receptors, CD95/APO-1 (Fas), and the TNF receptors. Cd can alter the CD95/APO-1 (Fas)/Fas ligand (FasL) signaling pathway, especially in neuronal cells. Moreover, it can increase inflammation markers, including TNF-ɑ and NF-kB, in nephrotic cells, leading to apoptosis by caspase-3 activation. On the other hand, Cd can also influence the intrinsic pathway of apoptosis, increasing the expression of the pro-apoptotic protein, Bax, and suppressing the anti-apoptotic protein, Bcl-2 [84]. In this case, a decrease in the Bcl-2/Bax ratio causes the release of cytochrome c from mitochondria and Ca from the endoplasmic reticulum (ER): this leads to the caspase cascade, with the final activation of caspase-3 and cell death. The activation of caspase-3 has a crucial role in cell death by Cd-mediated apoptosis, so it is logical to hypothesize that the inhibition of caspase-3 can prevent Cd toxicity. There are several studies in this field, claiming that certain natural substances such as taxifolin can protect from Cd apoptosis in skin cells by changing the activity of caspase-3 and -7, and other apoptotic factors [85]. Cd toxicity seems to also play a role in autophagy processes, even if this role has not been understood properly [86]. Cd shows, in fact, a different output for the cell fate in different cell lines. Cd-induced autophagy seems to promote apoptosis in skin epidermal cells [87], in mouse spleen, in neuronal cells [88], in mouse renal tubular epithelial cells, and in rat proximal tubular epithelial cells [89]. On the other hand, Cd-exposed human bronchial epithelial cells (BEAS-2BR) appeared to be autophagy-deficient, down-streaming the anti-apoptotic proteins Bcl-2 and Bcl-xl, apoptosis resistance, and possible carcinogenesis [90].
Cd alters the epigenetic signatures in the DNA of the placenta and of newborns, and some studies indicated marked sexual differences for Cd-related DNA methylation changes. Associations between early Cd exposure and DNA methylation might reflect interference with de novo DNA methyltransferases. In particular, the association of cadmium with the DNA methylation of certain CpG sites within the genes of interest in organ development, glucocorticoid synthesis, and cell death has been reported [91,92].

5.3. Cd-Metalloproteins with Relevance to Carcinogenesis

According to a recent study, the Cd-metalloproteome might constitute up to 18.4% of the entire human proteome [93]. In the present study, we identified which genes encoding Cd-metalloproteome are genetic markers for 12 cancer types (urothelial bladder carcinoma, breast cancer, colorectal adenocarcinoma, glioblastoma, head and neck squamous cell carcinoma, kidney cancer, acute myeloid leukemia, lung adenocarcinoma, lung squamous cell carcinoma, ovarian cancer, uterine corpus endometrial carcinoma, and multiple myeloma) using the Cancer Genome Atlas (TCGA) pan-cancer repository [94]. Specifically, 12 Cd-binding proteins have been characterized as cancer-related, and their names, abundance, and tissue-specific expression in the human body are listed in Table 1. Figure 5 shows the 12 cancer types associated with each specific gene-encoding Cd-metalloproteome.
Five of these proteins (tyrosine-protein kinase Blk, polypeptide N-acetylgalactosaminyltransferase 10, tyrosine-protein kinase Lck, cAMP-dependent protein kinase catalytic subunit alpha, and TGF-beta receptor type-2) are transferases. B-lymphoid tyrosine kinase (Blk) is an oncogene and a potential target for therapy with dasatinib in cutaneous T-cell lymphoma (CTCL) [95]. Gene knockdown experiments showed that Blk promoted the proliferation of malignant T-cells from CTCL patients, suggesting that Blk may function as an oncogene. Blk is also implicated in childhood acute lymphoblastic leukemia [96].
N-acetylgalactosaminyltransferase genes (GALNTs) and proteins (GalNAcTs) are involved in cancer biology. Aberrant O-glycosylation by GalNAcTs activates a wide range of proteins that carry out the interactions of sessile and motile cells affecting organogenesis and the responses to agonists, stimulating the hyperproliferation and metastization of neoplastic cells [97].
cAMP-dependent protein kinase catalytic subunit alpha is a multi-unit protein kinase that mediates the signal transduction of G-protein-coupled receptors through its activation upon cAMP binding. The cAMP/PKA signaling pathway is altered in different cancers and may be exploited for cancer therapy and/or diagnosis via cell cycle regulation and stimulated cell growth [98].
TGF-beta receptor type-2 TGF (TGFBR2) is the ligand-binding receptor for all members of the TGF-β family. TGF-beta receptor type-2 expression in cancer-associated fibroblasts regulates breast cancer cell growth and survival and is a prognostic marker in pre-menopausal breast cancer [99]. The mutational inactivation of TGFBR2 is the most common genetic event affecting the TGF-β signaling pathway and occurs in ∼20–30% of all colon cancers [100].
Among the above-mentioned 12 Cd-binding proteins, the most abundant, based on the protein abundance database PaxDB (https://pax-db.org/ (accessed on 21 October 2022)) [101], is heat shock protein HSP 90-beta, which is ranked in the top 5% of human proteins (Figure S1 in the Supplementary Materials). Heat-shock proteins are found at increased levels in many solid tumors and hematological malignancies. Their expression may, in part, account for the ability of malignant cells to maintain protein homeostasis, even in the hostile hypoxic and acidotic microenvironment of the tumor. HSP 90-beta is known as a cancer chaperone, required for the stability and function of multiple mutated, chimeric, and/or over-expressed signaling proteins that promote cancer cell growth and/or survival. It has also been implicated in many other crucial steps of carcinogenesis: the inhibition of programmed cell death and replicative senescence, the induction of tumor angiogenesis, and the activation of invasion and metastasis [102,103,104,105].
A variety of Hsp90 inhibitors have shown antitumor effects as a single agent and in combination with chemotherapy. Current Hsp90 inhibitors are categorized into several classes, based on the distinct modes of inhibition, including (i) the blockade of ATP binding, (ii) the disruption of cochaperone/Hsp90 interactions, (iii) the antagonism of client/Hsp90 associations, and (iv) interference with the post-translational modifications of Hsp90 [106].

6. Environmental Remediation of Cadmium

The very first prevention step regarding Cd pollution (which is actually one of the main causes of toxicity for humans) is the removal of Cd from contaminated soil and water. This task can be achieved using both physical and chemical methods: on the one hand, there are physical methods, such as adsorption, ion exchange, and reverse osmosis, while, on the other hand, there are chemical methods, which involve precipitation, electrolysis, and solvent extraction [107]. Besides these classical methods, natural remediation from plants and other organisms has also been investigated. As an environmental pollutant, Cd, as with some other metals, is different from organic pollutants, which can be degraded by microorganisms. Microbial remediation, based on metal biosorption, is currently considered to be an efficient strategy for the detoxification of Cd from contaminated waste, water, soil, and sediments [108,109]. In the past few years, the literature has revealed an outstanding variety of microorganisms that are able to process and eliminate Cd from the environment, such as bacteria, fungi, yeasts, and different kinds of algae. In particular, bacteria such as the genera Aeromonas, Bacillus, and Pseudomonas, and diatom microalgae are able to efficiently accumulate Cd [110,111].
All these methods can be effective in a controlled environment, but they all need very specific conditions of climate, temperature, and humidity to be carried out in a specific ecosystem [107]. More recently, a different remediation method called “phytoremediation” has been proposed [107,112]. In the literature, it is possible to find a huge number of plant species with the peculiar skill of Cd accumulation; these plants are called Cd hyperaccumulators [113].

7. Detoxifying Agents and Chelating Agents for the Prevention and Treatment of Cadmium Toxicity

The toxicity of Cd has also been related to its neurotoxic effects, which develop via the changes that Cd induces in the brain enzymes. Trace elements such as Zn and selenium (Se) have been investigated as preventive agents against Cd toxicity, because of their functional role in the brain. As discussed above, Zn2+ possesses chemical and physical characteristics that are similar to Cd2+ and naturally compete with it for the binding sites in enzymes. Zn also induces the synthesis of the CNS-specific MT III [114], which has a high affinity for Cd and can cause detoxification by binding it. Selenium is recognized to be effective in improving antioxidant defense, immune functions, and metabolic homeostasis, with a critical role in anti-aging [115]. It plays a role in the depletion of Cd from the body by protecting it against Cd toxicity in a number of different organs, including the brain [116]. Se species combine with Cd ions, and both are excreted out of the body via the bile system. Therefore, there is less Se to form GSH peroxidase, one of the body’s main antioxidants [12]. This results in the formation of greater levels of ROS and hydrogen peroxide, with relative cellular damage. Se supplementation will, therefore, be useful for increasing immunity and effectively restoring the body’s antioxidant defense [115]. Se is also a cofactor of GPx, an antioxidant enzyme that can contribute to reducing Cd oxidative effects [117]. Consequently, a preventive approach to counteract chronic Cd exposure is based on the use of trace element supplements as detoxifying agents. Zn and Mg supplementation also has the potential to modulate and mitigate Cd intoxication in several organs [26,27]. In addition, several studies in animal models showed that vitamins A, C, E are able to decrease the toxic effects of Cd in the kidney, liver, spleen, blood, bones, and brain. However, further human studies are needed to clarify the role of these antioxidant vitamins in reducing Cd-induced toxicity [118].
Another strategy of prevention is action at the adsorption level, in the gastrointestinal tract. Research has indeed demonstrated that some strains of lactic acid bacteria (LAB) may be able to bind and remove heavy metals, such as Cd and lead (Pb) [119]. Moreover, probiotics such as these can have antioxidant properties for the human body, being effective against Cd-induced oxidative stress [120]. One of the most studied and convincing probiotics in this field is the Lactobacillus plantarum CCFM8610, which was demonstrated to have good Cd-binding ability and to be capable of protecting the liver and kidneys of mice in acute Cd intoxication [121]. This particular effect is due to the Cd sequestration at the intestinal level, reducing the bioavailability of the metal ion in tissues, and decreasing Cd-induced oxidative stress [122]. It has been demonstrated that L. plantarum also has a direct effect of protection against Cd-induced oxidative stress, increasing the MT protein levels in the liver [123]. The role of N-acetyl-cysteine (NAC) has also been investigated in the past few years as a cysteine group donor with a possible key role in the biosynthesis of MTs. It has been discovered that exogenous NAC can increase the production of MTs and enhance the possibility of binding and eliminating Cd2+ [124]. Cadmium intoxication from poisoning is extremely uncommon. The literature reported a case of Cd intoxication solved with the concomitant administration of GSH, along with the chelating agent, Ca-EDTA [125]. However, EDTA could increase the Cd levels in the kidneys, leading to renal dysfunction [126], as is the case when using dimercaprol (BAL) [127]. Penicillamine (DPA) has been found not to be efficient in Cd overdoses [128]. Meso-2,3-dimercaptosuccinic acid (succimer, DMSA) is a metal chelator not able to reach the intracellular Cd that is bound to MTs and is stored in the liver and kidneys [129]. However, a water-soluble, lipophilic chelating agent, MiADMSA, a C5-branched alkyl monoester of DMSA, can reach the stored Cd intracellularly, as well as other DMSA derivatives such as Monomethyl DMSA (MmDMSA) and Monocyclohexyl DMSA (MchDMSA) [130]. It has also been reported that a proper combination of selected chelating agents can be considered as more effective than mono-therapy [131]. This strategy considers the concomitant use of DMSA and MiADMSA, or calcium trisodium diethylene triamine pentaacetate (CaDTPA), or NAC [131].

8. Conclusions

In summary, this review describes the research associated with human exposure to cadmium, with an emphasis on its biological targets, toxic effects (mainly carcinogenic), and therapeutic approaches. The toxicity of Cd2+ is, in part, due to its ability to mimic Zn2+ and Ca2+ in their biological roles. The presence of metallothioneins protects humans from chronic exposure to low concentrations of Cd2+. The mechanisms of Cd-induced carcinogenesis have been related to oxidative stress with the inhibition of antioxidant enzymes, the promotion of lipid peroxidation, and interference with DNA repair systems. Strategies of prevention and treatment of Cd toxicity include the administering of trace elements, e.g., Se, as detoxifying agents, therapeutic schemes involving the use of antioxidant vitamins (A, C, E), and action at adsorption level in the gastrointestinal tract, by means of several probiotics.
Future research directions concerning the topic of this review are predicted to be: (i) an investigation of the important role of other metal-binding proteins, except the metallothioneins, and/or new receptors involved in the binding and transport of Cd2+; (ii) advanced molecular profiling of the events associated with Cd carcinogenesis in model systems, which may allow the development of expression signatures for Cd-induced cancers; (iii) elucidation of the mechanisms for Cd2+ clustering in the cysteine-rich regions of metallothioneins; (iv) clarification of the mechanisms that are responsible for “phytoremediation”; (v) further studies on the role played by Se in the depletion of Cd from the body; and (vi) development of selective chelating agents for Cd2+ removal from aqueous environments via solvent extraction, an area in which our research groups are currently involved [132,133].

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/biom13010036/s1, Figure S1: Protein abundance histogram of human proteins; Table S1: Percentage distribution of geometries for Cd2+ complexed with proteins according to RCSB PDB database; Table S2: Percentage distribution of metal ions substituted by Cd2+ according to the RCSB PDB database.

Author Contributions

Conceptualization, M.P. and A.P.; writing—original draft preparation, M.P., A.P. and C.T.C.; writing—review and editing, M.P., A.P., C.T.C., S.P.P., V.B., S.M. and M.A.Z.; visualization M.P., A.P. and C.T.C.; M.P., V.B., S.P.P., S.M. and M.A.Z. contributed to the final version of the manuscript. Revision: M.P. and S.P.P. M.P. supervised the project. All authors have read and agreed to the published version of the manuscript.

Funding

S.M. and M.A.Z. acknowledge Università degli Studi di Sassari (UNISS) for the financial support received within the program “Fondo di Ateneo per la ricerca 2019, FAR 2019” (Rep. 2467, Prot. 94737 07/08/2019), and M.P. within the program “Fondo di Ateneo per la ricerca 2020, FAR 2020” (Rep. 2465, Prot. 0097985 01/09/2020).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All data generated or analyzed during this study are included in this article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Sharma, H.; Rawal, N.; Mathew, B.B. The characteristics, toxicity and effects of cadmium. Int. J. Nanotechnol. 2015, 3, 1–9. [Google Scholar]
  2. Zoroddu, M.A.; Aaseth, J.; Crisponi, G.; Medici, S.; Peana, M.; Nurchi, V.M. The essential metals for humans: A brief overview. J. Inorg. Biochem. 2019, 195, 120–129. [Google Scholar] [CrossRef] [PubMed]
  3. Maret, W.; Moulis, J.M. The bioinorganic chemistry of cadmium in the context of its toxicity. In Metal ions in life sciences; Springer: Berlin/Heidelberg, Germany, 2013; Volume 11, pp. 1–29. [Google Scholar]
  4. Mason, R.P. Trace Metals in Aquatic Systems, 1st ed.; Wiley-Blackwell: Hoboken, NJ, USA, 2013. [Google Scholar]
  5. Nordberg, M.; Nordberg, G.F. Metallothionein and Cadmium Toxicology-Historical Review and Commentary. Biomolecules 2022, 12, 360. [Google Scholar] [CrossRef] [PubMed]
  6. Krezel, A.; Maret, W. The Bioinorganic Chemistry of Mammalian Metallothioneins. Chem. Rev. 2021, 121, 14594–14648. [Google Scholar] [CrossRef] [PubMed]
  7. Klaassen, C.D.; Liu, J.; Diwan, B.A. Metallothionein protection of cadmium toxicity. Toxicol. Appl. Pharmacol. 2009, 238, 215–220. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Egli, D.; Domenech, J.; Selvaraj, A.; Balamurugan, K.; Hua, H.; Capdevila, M.; Georgiev, O.; Schaffner, W.; Atrian, S. The four members of the Drosophila metallothionein family exhibit distinct yet overlapping roles in heavy metal homeostasis and detoxification. Genes Cells 2006, 11, 647–658. [Google Scholar] [CrossRef]
  9. Sinicropi, M.S.; Amantea, D.; Caruso, A.; Saturnino, C. Chemical and biological properties of toxic metals and use of chelating agents for the pharmacological treatment of metal poisoning. Arch. Toxicol. 2010, 84, 501–520. [Google Scholar] [CrossRef]
  10. Bernhoft, R.A. Cadmium toxicity and treatment. Sci. World J. 2013, 2013, 394652. [Google Scholar] [CrossRef]
  11. Jarup, L.; Rogenfelt, A.; Elinder, C.G.; Nogawa, K.; Kjellstrom, T. Biological half-time of cadmium in the blood of workers after cessation of exposure. Scand J. Work Environ. Health 1983, 9, 327–331. [Google Scholar] [CrossRef]
  12. Rani, A.; Kumar, A.; Lal, A.; Pant, M. Cellular mechanisms of cadmium-induced toxicity: A review. Int. J. Environ. Health Res. 2014, 24, 378–399. [Google Scholar] [CrossRef]
  13. Branca, J.J.V.; Pacini, A.; Gulisano, M.; Taddei, N.; Fiorillo, C.; Becatti, M. Cadmium-Induced Cytotoxicity: Effects on Mitochondrial Electron Transport Chain. Front. Cell Dev. Biol. 2020, 8, 604377. [Google Scholar] [CrossRef] [PubMed]
  14. Gasmi, A.; Peana, M.; Arshad, M.; Butnariu, M.; Menzel, A.; Bjørklund, G. Krebs cycle: Activators, inhibitors and their roles in the modulation of carcinogenesis. Arch. Toxicol. 2021, 95, 1161–1178. [Google Scholar] [CrossRef] [PubMed]
  15. Koedrith, P.; Seo, Y.R. Advances in carcinogenic metal toxicity and potential molecular markers. Int. J. Mol. Sci. 2011, 12, 9576–9595. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Beyersmann, D.; Hartwig, A. Carcinogenic metal compounds: Recent insight into molecular and cellular mechanisms. Arch. Toxicol. 2008, 82, 493. [Google Scholar] [CrossRef] [PubMed]
  17. Bjørklund, G.; Zou, L.; Peana, M.; Chasapis, C.T.; Hangan, T.; Lu, J.; Maes, M. The Role of the Thioredoxin System in Brain Diseases. Antioxidants 2022, 11, 2161. [Google Scholar] [CrossRef]
  18. Pearson, R.G. Chemical hardness and density functional theory. J. Chem. Sci. 2005, 117, 369–377. [Google Scholar] [CrossRef]
  19. Peana, M.; Pelucelli, A.; Medici, S.; Cappai, R.; Nurchi, V.M.; Zoroddu, M.A. Metal Toxicity and Speciation: A Review. Curr. Med. Chem. 2021, 28, 7190–7208. [Google Scholar] [CrossRef]
  20. Remelli, M.; Nurchi, V.M.; Lachowicz, J.I.; Medici, S.; Zoroddu, M.A.; Peana, M. Competition between Cd(II) and other divalent transition metal ions during complex formation with amino acids, peptides, and chelating agents. Coord. Chem. Rev. 2016, 327-328, 55–69. [Google Scholar] [CrossRef]
  21. Wernimont, A.K.; Huffman, D.L.; Lamb, A.L.; O’Halloran, T.V.; Rosenzweig, A.C. Structural basis for copper transfer by the metallochaperone for the Menkes/Wilson disease proteins. Nat. Struct. Biol. 2000, 7, 766–771. [Google Scholar] [CrossRef]
  22. Muramoto, K.; Hirata, K.; Shinzawa-Itoh, K.; Yoko-o, S.; Yamashita, E.; Aoyama, H.; Tsukihara, T.; Yoshikawa, S. A histidine residue acting as a controlling site for dioxygen reduction and proton pumping by cytochrome <em>c</em> oxidase. Proc. Natl. Acad. Sci. USA 2007, 104, 7881. [Google Scholar] [CrossRef] [Green Version]
  23. Zangger, K.; Oz, G.; Otvos, J.D.; Armitage, I.M. Three-dimensional solution structure of mouse [Cd7]-metallothionein-1 by homonuclear and heteronuclear NMR spectroscopy. Protein Sci. 1999, 8, 2630–2638. [Google Scholar] [CrossRef] [PubMed]
  24. Putignano, V.; Rosato, A.; Banci, L.; Andreini, C. MetalPDB in 2018: A database of metal sites in biological macromolecular structures. Nucleic Acids Res. 2018, 46, D459–D464. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Valasatava, Y.; Rosato, A.; Cavallaro, G.; Andreini, C. MetalS(3), a database-mining tool for the identification of structurally similar metal sites. J. Biol. Inorg. Chem. 2014, 19, 937–945. [Google Scholar] [CrossRef] [PubMed]
  26. Bulat, Z.; Dukic-Cosic, D.; Antonijevic, B.; Bulat, P.; Vujanovic, D.; Buha, A.; Matovic, V. Effect of magnesium supplementation on the distribution patterns of zinc, copper, and magnesium in rabbits exposed to prolonged cadmium intoxication. Sci. World J. 2012, 2012, 572514. [Google Scholar] [CrossRef] [Green Version]
  27. Bulat, Z.P.; Djukic-Cosic, D.; Malicevic, Z.; Bulat, P.; Matovic, V. Zinc or magnesium supplementation modulates cd intoxication in blood, kidney, spleen, and bone of rabbits. Biol. Trace Elem. Res. 2008, 124, 110–117. [Google Scholar] [CrossRef]
  28. Bekiari, V.; Lianos, P. High-Yield Luminescence from Cadmium Sulfide Nanoclusters Supported in a Poly(ethylene glycol) Oligomer. Langmuir 2000, 16, 3561–3563. [Google Scholar] [CrossRef]
  29. Sobiech, M.; Bujak, P.; Luliński, P.; Pron, A. Semiconductor nanocrystal–polymer hybrid nanomaterials and their application in molecular imprinting. Nanoscale 2019, 11, 12030–12074. [Google Scholar] [CrossRef]
  30. European Parliament. Directive (EU) 2017/2102 of the European Parliament and of the Council of 15 November 2017 amending Directive 2011/65/EU on the restriction of the use of certain hazardous substances in electrical and electronic equipment (Text with EEA relevance.). Off. J. Eur. Union 2017, L305. [Google Scholar]
  31. Xu, L.; Huo, X.; Liu, Y.; Zhang, Y.; Qin, Q.; Xu, X. Hearing loss risk and DNA methylation signatures in preschool children following lead and cadmium exposure from an electronic waste recycling area. Chemosphere 2020, 246, 125829. [Google Scholar] [CrossRef]
  32. Baloch, S.; Kazi, T.G.; Baig, J.A.; Afridi, H.I.; Arain, M.B. Occupational exposure of lead and cadmium on adolescent and adult workers of battery recycling and welding workshops: Adverse impact on health. Sci. Total. Environ. 2020, 720, 137549. [Google Scholar] [CrossRef] [PubMed]
  33. Hayat, M.T.; Nauman, M.; Nazir, N.; Ali, S.; Bangash, N. Chapter 7—Environmental Hazards of Cadmium: Past, Present, and Future. In Cadmium Toxicity and Tolerance in Plants; Hasanuzzaman, M., Prasad, M.N.V., Fujita, M., Eds.; Academic Press: Cambridge, MA, USA, 2019; pp. 163–183. [Google Scholar]
  34. Chavez, E.; He, Z.L.; Stoffella, P.J.; Mylavarapu, R.S.; Li, Y.C.; Moyano, B.; Baligar, V.C. Concentration of cadmium in cacao beans and its relationship with soil cadmium in southern Ecuador. Sci. Total. Environ. 2015, 533, 205–214. [Google Scholar] [CrossRef] [PubMed]
  35. Kim, K.; Melough, M.M.; Vance, T.M.; Noh, H.; Koo, S.I.; Chun, O.K. Dietary Cadmium Intake and Sources in the US. Nutrients 2018, 11. [Google Scholar] [CrossRef] [PubMed]
  36. Watanabe, T.; Zhang, Z.W.; Moon, C.S.; Shimbo, S.; Nakatsuka, H.; Matsuda-Inoguchi, N.; Higashikawa, K.; Ikeda, M. Cadmium exposure of women in general populations in Japan during 1991-1997 compared with 1977-1981. Int. Arch. Occup. Environ. Health 2000, 73, 26–34. [Google Scholar] [CrossRef] [PubMed]
  37. Song, Y.; Wang, Y.; Mao, W.; Sui, H.; Yong, L.; Yang, D.; Jiang, D.; Zhang, L.; Gong, Y. Dietary cadmium exposure assessment among the Chinese population. PLoS ONE 2017, 12, e0177978. [Google Scholar] [CrossRef] [Green Version]
  38. Satarug, S.; Baker, J.R.; Urbenjapol, S.; Haswell-Elkins, M.; Reilly, P.E.; Williams, D.J.; Moore, M.R. A global perspective on cadmium pollution and toxicity in non-occupationally exposed population. Toxicol. Lett. 2003, 137, 65–83. [Google Scholar] [CrossRef]
  39. Horiguchi, H.; Oguma, E.; Sasaki, S.; Miyamoto, K.; Ikeda, Y.; Machida, M.; Kayama, F. Dietary exposure to cadmium at close to the current provisional tolerable weekly intake does not affect renal function among female Japanese farmers. Environ. Res. 2004, 95, 20–31. [Google Scholar] [CrossRef]
  40. Nogawa, K.; Suwazono, Y.; Nishijo, M.; Sakurai, M.; Ishizaki, M.; Morikawa, Y.; Watanabe, Y.; Kido, T.; Nakagawa, H. Increase of lifetime cadmium intake dose-dependently increased all cause of mortality in female inhabitants of the cadmium-polluted Jinzu River basin, Toyama, Japan. Environ. Res. 2018, 164, 379–384. [Google Scholar] [CrossRef]
  41. Nishijo, M.; Nogawa, K.; Suwazono, Y.; Kido, T.; Sakurai, M.; Nakagawa, H. Lifetime Cadmium Exposure and Mortality for Renal Diseases in Residents of the Cadmium-Polluted Kakehashi River Basin in Japan. Toxics 2020, 8. [Google Scholar] [CrossRef]
  42. Uetani, M.; Kobayashi, E.; Suwazono, Y.; Kido, T.; Nogawa, K. Cadmium exposure aggravates mortality more in women than in men. Int. J. Environ. Health Res. 2006, 16, 273–279. [Google Scholar] [CrossRef]
  43. Duggan, R.E.; Lipscomb, G.Q. Dietary intake of pesticide chicals in the United States (II), June 1966—April 1968. Pestic. Monit. J. 1969, 2, 153–162. [Google Scholar]
  44. Choudhury, H.; Harvey, T.; Thayer, W.C.; Lockwood, T.F.; Stiteler, W.M.; Goodrum, P.E.; Hassett, J.M.; Diamond, G.L. Urinary cadmium elimination as a biomarker of exposure for evaluating a cadmium dietary exposure--biokinetics model. J. Toxicol. Environ. Health A 2001, 63, 321–350. [Google Scholar] [CrossRef] [PubMed]
  45. Mahjoub, M.; Fadlaoui, S.; El Maadoudi, M.; Smiri, Y. Mercury, Lead, and Cadmium in the Muscles of Five Fish Species from the Mechraa-Hammadi Dam in Morocco and Health Risks for Their Consumers. J. Toxicol. 2021, 2021, 8865869. [Google Scholar] [CrossRef] [PubMed]
  46. Mok, J.S.; Yoo, H.D.; Kim, P.H.; Yoon, H.D.; Park, Y.C.; Lee, T.S.; Kwon, J.Y.; Son, K.T.; Lee, H.J.; Ha, K.S.; et al. Bioaccumulation of heavy metals in oysters from the southern coast of Korea: Assessment of potential risk to human health. Bull. Environ. Contam. Toxicol. 2015, 94, 749–755. [Google Scholar] [CrossRef] [PubMed]
  47. Mok, J.S.; Kwon, J.Y.; Son, K.T.; Choi, W.S.; Shim, K.B.; Lee, T.S.; Kim, J.H. Distribution of heavy metals in muscles and internal organs of Korean cephalopods and crustaceans: Risk assessment for human health. J. Food Prot. 2014, 77, 2168–2175. [Google Scholar] [CrossRef] [PubMed]
  48. Shi, W.; Zhao, X.; Han, Y.; Che, Z.; Chai, X.; Liu, G. Ocean acidification increases cadmium accumulation in marine bivalves: A potential threat to seafood safety. Sci. Rep. 2016, 6, 20197. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Conterato, G.M.; Bulcao, R.P.; Sobieski, R.; Moro, A.M.; Charao, M.F.; de Freitas, F.A.; de Almeida, F.L.; Moreira, A.P.; Roehrs, M.; Tonello, R.; et al. Blood thioredoxin reductase activity, oxidative stress and hematological parameters in painters and battery workers: Relationship with lead and cadmium levels in blood. J. Appl. Toxicol. 2013, 33, 142–150. [Google Scholar] [CrossRef]
  50. Oberdorster, G. Airborne cadmium and carcinogenesis of the respiratory tract. Scand J. Work Environ. Health 1986, 12, 523–537. [Google Scholar] [CrossRef]
  51. Mannino, D.M.; Holguin, F.; Greves, H.M.; Savage-Brown, A.; Stock, A.L.; Jones, R.L. Urinary cadmium levels predict lower lung function in current and former smokers: Data from the Third National Health and Nutrition Examination Survey. Thorax 2004, 59, 194–198. [Google Scholar] [CrossRef] [Green Version]
  52. Satarug, S.; Moore, M.R. Adverse health effects of chronic exposure to low-level cadmium in foodstuffs and cigarette smoke. Environ. Health Perspect 2004, 112, 1099–1103. [Google Scholar] [CrossRef] [Green Version]
  53. Barregard, L.; Fabricius-Lagging, E.; Lundh, T.; Molne, J.; Wallin, M.; Olausson, M.; Modigh, C.; Sallsten, G. Cadmium, mercury, and lead in kidney cortex of living kidney donors: Impact of different exposure sources. Environ. Res. 2010, 110, 47–54. [Google Scholar] [CrossRef]
  54. Bjørklund, G.; Oliinyk, P.; Lysiuk, R.; Rahaman, M.S.; Antonyak, H.; Lozynska, I.; Lenchyk, L.; Peana, M. Arsenic intoxication: General aspects and chelating agents. Arch. Toxicol. 2020, 94, 1879–1897. [Google Scholar] [CrossRef] [PubMed]
  55. Liu, H.; Gao, H.; Long, M.; Fu, H.; Alvarez, P.J.J.; Li, Q.; Zheng, S.; Qu, X.; Zhu, D. Sunlight Promotes Fast Release of Hazardous Cadmium from Widely-Used Commercial Cadmium Pigment. Environ. Sci. Technol. 2017, 51, 6877–6886. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Chavatte, L.; Juan, M.; Mounicou, S.; Leblanc Noblesse, E.; Pays, K.; Nizard, C.; Bulteau, A.L. Elemental and molecular imaging of human full thickness skin after exposure to heavy metals. Metallomics 2020, 12, 1555–1562. [Google Scholar] [CrossRef] [PubMed]
  57. Akesson, A.; Lundh, T.; Vahter, M.; Bjellerup, P.; Lidfeldt, J.; Nerbrand, C.; Samsioe, G.; Stromberg, U.; Skerfving, S. Tubular and glomerular kidney effects in Swedish women with low environmental cadmium exposure. Environ. Health Perspect 2005, 113, 1627–1631. [Google Scholar] [CrossRef] [Green Version]
  58. Kjellstrom, T. Mechanism and epidemiology of bone effects of cadmium. IARC Sci. Publ. 1992, 301–310. [Google Scholar]
  59. Pizent, A.; Tariba, B.; Zivkovic, T. Reproductive toxicity of metals in men. Arh. Hig. Rada Toksikol. 2012, 63 (Suppl. 1), 35–46. [Google Scholar] [CrossRef]
  60. Thompson, J.; Bannigan, J. Cadmium: Toxic effects on the reproductive system and the embryo. Reprod. Toxicol. 2008, 25, 304–315. [Google Scholar] [CrossRef]
  61. Eum, K.D.; Lee, M.S.; Paek, D. Cadmium in blood and hypertension. Sci. Total Environ. 2008, 407, 147–153. [Google Scholar] [CrossRef]
  62. Gallagher, C.M.; Meliker, J.R. Blood and urine cadmium, blood pressure, and hypertension: A systematic review and meta-analysis. Environ. Health Perspect 2010, 118, 1676–1684. [Google Scholar] [CrossRef] [Green Version]
  63. Edwards, J.R.; Prozialeck, W.C. Cadmium, diabetes and chronic kidney disease. Toxicol. Appl. Pharmacol. 2009, 238, 289–293. [Google Scholar] [CrossRef] [Green Version]
  64. Horiguchi, H.; Teranishi, H.; Niiya, K.; Aoshima, K.; Katoh, T.; Sakuragawa, N.; Kasuya, M. Hypoproduction of erythropoietin contributes to anemia in chronic cadmium intoxication: Clinical study on Itai-itai disease in Japan. Arch. Toxicol. 1994, 68, 632–636. [Google Scholar] [CrossRef] [PubMed]
  65. Fujiwara, Y.; Lee, J.Y.; Banno, H.; Imai, S.; Tokumoto, M.; Hasegawa, T.; Seko, Y.; Nagase, H.; Satoh, M. Cadmium induces iron deficiency anemia through the suppression of iron transport in the duodenum. Toxicol. Lett. 2020, 332, 130–139. [Google Scholar] [CrossRef] [PubMed]
  66. Li, F.J.; Surolia, R.; Singh, P.; Dsouza, K.G.; Stephens, C.T.; Wang, Z.; Liu, R.M.; Bae, S.; Kim, Y.I.; Athar, M.; et al. Fibrinogen mediates cadmium-induced macrophage activation and serves as a predictor of cadmium exposure in chronic obstructive pulmonary disease. Am. J. Physiol. Lung Cell Mol. Physiol. 2022, 322, L593–L606. [Google Scholar] [CrossRef] [PubMed]
  67. Li, F.J.; Surolia, R.; Li, H.; Wang, Z.; Liu, G.; Liu, R.M.; Mirov, S.B.; Athar, M.; Thannickal, V.J.; Antony, V.B. Low-dose cadmium exposure induces peribronchiolar fibrosis through site-specific phosphorylation of vimentin. Am. J. Physiol. Lung Cell Mol. Physiol. 2017, 313, L80–L91. [Google Scholar] [CrossRef]
  68. Rinaldi, M.; Micali, A.; Marini, H.; Adamo, E.B.; Puzzolo, D.; Pisani, A.; Trichilo, V.; Altavilla, D.; Squadrito, F.; Minutoli, L. Cadmium, Organ Toxicity and Therapeutic Approaches: A Review on Brain, Kidney and Testis Damage. Curr. Med. Chem. 2017, 24, 3879–3893. [Google Scholar] [CrossRef]
  69. Branca, J.J.V.; Morucci, G.; Pacini, A. Cadmium-induced neurotoxicity: Still much ado. Neural. Regen. Res. 2018, 13, 1879–1882. [Google Scholar] [CrossRef]
  70. Filipic, M.; Fatur, T.; Vudrag, M. Molecular mechanisms of cadmium induced mutagenicity. Hum. Exp. Toxicol. 2006, 25, 67–77. [Google Scholar] [CrossRef]
  71. Joseph, P.; Muchnok, T.K.; Klishis, M.L.; Roberts, J.R.; Antonini, J.M.; Whong, W.-Z.; Ong, T.-m. Cadmium-Induced Cell Transformation and Tumorigenesis Are Associated with Transcriptional Activation of c-fos, c-jun, and c-myc Proto-Oncogenes: Role of Cellular Calcium and Reactive Oxygen Species. Toxicol. Sci. 2001, 61, 295–303. [Google Scholar] [CrossRef] [Green Version]
  72. Liu, J.; Qu, W.; Kadiiska, M.B. Role of oxidative stress in cadmium toxicity and carcinogenesis. Toxicol. Appl. Pharmacol. 2009, 238, 209–214. [Google Scholar] [CrossRef] [Green Version]
  73. Ogasawara, Y.; Takeda, Y.; Takayama, H.; Nishimoto, S.; Ichikawa, K.; Ueki, M.; Suzuki, T.; Ishii, K. Significance of the rapid increase in GSH levels in the protective response to cadmium exposure through phosphorylated Nrf2 signaling in Jurkat T-cells. Free Radic. Biol. Med. 2014, 69, 58–66. [Google Scholar] [CrossRef]
  74. Schwerdtle, T.; Ebert, F.; Thuy, C.; Richter, C.; Mullenders, L.H.F.; Hartwig, A. Genotoxicity of Soluble and Particulate Cadmium Compounds: Impact on Oxidative DNA Damage and Nucleotide Excision Repair. Chem. Res. Toxicol. 2010, 23, 432–442. [Google Scholar] [CrossRef] [PubMed]
  75. Birkou, M.; Chasapis, C.T.; Marousis, K.D.; Loutsidou, A.K.; Bentrop, D.; Lelli, M.; Herrmann, T.; Carthy, J.M.; Episkopou, V.; Spyroulias, G.A. A Residue Specific Insight into the Arkadia E3 Ubiquitin Ligase Activity and Conformational Plasticity. J. Mol. Biol. 2017, 429, 2373–2386. [Google Scholar] [CrossRef] [PubMed]
  76. Chasapis, C.T.; Kandias, N.G.; Episkopou, V.; Bentrop, D.; Spyroulias, G.A. NMR-based insights into the conformational and interaction properties of Arkadia RING-H2 E3 Ub ligase. Proteins Struct. Funct. Bioinform. 2012, 80, 1484–1489. [Google Scholar] [CrossRef] [PubMed]
  77. Fatur, T.; Lah, T.T.; Filipic, M. Cadmium inhibits repair of UV-, methyl methanesulfonate- and N-methyl-N-nitrosourea-induced DNA damage in Chinese hamster ovary cells. Mutat. Res. 2003, 529, 109–116. [Google Scholar] [CrossRef] [PubMed]
  78. Hsieh, P.; Yamane, K. DNA mismatch repair: Molecular mechanism, cancer, and ageing. Mech. Ageing Dev. 2008, 129, 391–407. [Google Scholar] [CrossRef] [Green Version]
  79. Jin, Y.H.; Clark, A.B.; Slebos, R.J.C.; Al-Refai, H.; Taylor, J.A.; Kunkel, T.A.; Resnick, M.A.; Gordenin, D.A. Cadmium is a mutagen that acts by inhibiting mismatch repair. Nat. Genet. 2003, 34, 326–329. [Google Scholar] [CrossRef]
  80. Wieland, M.; Levin, M.K.; Hingorani, K.S.; Biro, F.N.; Hingorani, M.M. Mechanism of Cadmium-Mediated Inhibition of Msh2-Msh6 Function in DNA Mismatch Repair. Biochemistry 2009, 48, 9492–9502. [Google Scholar] [CrossRef] [Green Version]
  81. Lützen, A.; Liberti, S.E.; Rasmussen, L.J. Cadmium inhibits human DNA mismatch repair in vivo. Biochem. Biophys. Res. Commun. 2004, 321, 21–25. [Google Scholar] [CrossRef]
  82. Wang, S.; Ren, X.; Hu, X.; Zhou, L.; Zhang, C.; Zhang, M. Cadmium-induced apoptosis through reactive oxygen species-mediated mitochondrial oxidative stress and the JNK signaling pathway in TM3 cells, a model of mouse Leydig cells. Toxicol. Appl. Pharmacol. 2019, 368, 37–48. [Google Scholar] [CrossRef]
  83. Yuan, Y.; Zhang, Y.; Zhao, S.; Chen, J.; Yang, J.; Wang, T.; Zou, H.; Wang, Y.; Gu, J.; Liu, X.; et al. Cadmium-induced apoptosis in neuronal cells is mediated by Fas/FasL-mediated mitochondrial apoptotic signaling pathway. Sci. Rep. 2018, 8, 8837. [Google Scholar] [CrossRef] [Green Version]
  84. Ghajari, H.; Hosseini, S.A.; Farsi, S. The Effect of Endurance Training Along with Cadmium Consumption on Bcl-2 and Bax Gene Expressions in Heart Tissue of Rats. Ann. Mil. Health Sci. Res. 2019, 17, e86795. [Google Scholar] [CrossRef]
  85. Moon, S.H.; Lee, C.M.; Nam, M.J. Cytoprotective effects of taxifolin against cadmium-induced apoptosis in human keratinocytes. Hum. Exp. Toxicol. 2019, 38, 992–1003. [Google Scholar] [CrossRef]
  86. Gu, J.; Wang, Y.; Liu, Y.; Shi, M.; Yin, L.; Hou, Y.; Zhou, Y.; Chu Wong, C.K.; Chen, D.; Guo, Z.; et al. Inhibition of Autophagy Alleviates Cadmium-Induced Mouse Spleen and Human B Cells Apoptosis. Toxicol. Sci. 2019, 170, 109–122. [Google Scholar] [CrossRef] [PubMed]
  87. Chiarelli, R.; Roccheri, M.C. Heavy Metals and Metalloids as Autophagy Inducing Agents: Focus on Cadmium and Arsenic. Cells 2012, 1. [Google Scholar] [CrossRef]
  88. Zhang, H.; Dong, X.; Zhao, R.; Zhang, R.; Xu, C.; Wang, X.; Liu, C.; Hu, X.; Huang, S.; Chen, L. Cadmium results in accumulation of autophagosomes-dependent apoptosis through activating Akt-impaired autophagic flux in neuronal cells. Cell. Signal. 2019, 55, 26–39. [Google Scholar] [CrossRef]
  89. Luo, T.; Zhang, H.; Yu, Q.; Liu, G.; Long, M.; Zhang, K.; Liu, W.; Song, R.; Bian, J.; Gu, J.; et al. ERK1/2 MAPK promotes autophagy to suppress ER stress-mediated apoptosis induced by cadmium in rat proximal tubular cells. Toxicol. Vitr. 2018, 52, 60–69. [Google Scholar] [CrossRef]
  90. Wang, Y.; Mandal, A.K.; Son, Y.-O.K.; Pratheeshkumar, P.; Wise, J.T.F.; Wang, L.; Zhang, Z.; Shi, X.; Chen, Z. Roles of ROS, Nrf2, and autophagy in cadmium-carcinogenesis and its prevention by sulforaphane. Toxicol. Appl. Pharmacol. 2018, 353, 23–30. [Google Scholar] [CrossRef] [PubMed]
  91. Vilahur, N.; Vahter, M.; Broberg, K. The Epigenetic Effects of Prenatal Cadmium Exposure. Curr. Environ. Health Rep. 2015, 2, 195–203. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Wang, B.; Li, Y.; Shao, C.; Tan, Y.; Cai, L. Cadmium and its epigenetic effects. Curr. Med. Chem. 2012, 19, 2611–2620. [Google Scholar] [CrossRef]
  93. Chasapis, C.T. Shared gene-network signatures between the human heavy metal proteome and neurological disorders and cancer types. Metallomics 2018, 10, 1678–1686. [Google Scholar] [CrossRef]
  94. Weinstein, J.N.; Collisson, E.A.; Mills, G.B.; Shaw, K.R.; Ozenberger, B.A.; Ellrott, K.; Shmulevich, I.; Sander, C.; Stuart, J.M. The Cancer Genome Atlas Pan-Cancer analysis project. Nat. Genet. 2013, 45, 1113–1120. [Google Scholar] [CrossRef] [PubMed]
  95. Petersen, D.L.; Krejsgaard, T.; Berthelsen, J.; Fredholm, S.; Willerslev-Olsen, A.; Sibbesen, N.A.; Bonefeld, C.M.; Andersen, M.H.; Francavilla, C.; Olsen, J.V.; et al. B-lymphoid tyrosine kinase (Blk) is an oncogene and a potential target for therapy with dasatinib in cutaneous T-cell lymphoma (CTCL). Leukemia 2014, 28, 2109–2112. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Montero-Ruiz, O.; Alcantara-Ortigoza, M.A.; Betancourt, M.; Juarez-Velazquez, R.; Gonzalez-Marquez, H.; Perez-Vera, P. Expression of RUNX1 isoforms and its target gene BLK in childhood acute lymphoblastic leukemia. Leuk Res. 2012, 36, 1105–1111. [Google Scholar] [CrossRef] [PubMed]
  97. Hussain, M.R.; Hoessli, D.C.; Fang, M. N-acetylgalactosaminyltransferases in cancer. Oncotarget 2016, 7, 54067–54081. [Google Scholar] [CrossRef] [PubMed]
  98. Sapio, L.; Di Maiolo, F.; Illiano, M.; Esposito, A.; Chiosi, E.; Spina, A.; Naviglio, S. Targeting protein kinase A in cancer therapy: An update. Excli. J. 2014, 13, 843–855. [Google Scholar]
  99. Busch, S.; Acar, A.; Magnusson, Y.; Gregersson, P.; Rydén, L.; Landberg, G. TGF-beta receptor type-2 expression in cancer-associated fibroblasts regulates breast cancer cell growth and survival and is a prognostic marker in pre-menopausal breast cancer. Oncogene 2015, 34, 27–38. [Google Scholar] [CrossRef]
  100. Biswas, S.; Chytil, A.; Washington, K.; Romero-Gallo, J.; Gorska, A.E.; Wirth, P.S.; Gautam, S.; Moses, H.L.; Grady, W.M. Transforming growth factor beta receptor type II inactivation promotes the establishment and progression of colon cancer. Cancer Res. 2004, 64, 4687–4692. [Google Scholar] [CrossRef] [Green Version]
  101. Wang, M.; Herrmann, C.J.; Simonovic, M.; Szklarczyk, D.; von Mering, C. Version 4.0 of PaxDb: Protein abundance data, integrated across model organisms, tissues, and cell-lines. PROTEOMICS 2015, 15, 3163–3168. [Google Scholar] [CrossRef] [Green Version]
  102. Den, R.B.; Lu, B. Heat shock protein 90 inhibition: Rationale and clinical potential. Ther. Adv. Med. Oncol. 2012, 4, 211–218. [Google Scholar] [CrossRef]
  103. Moser, C.; Lang, S.A.; Stoeltzing, O. Heat-shock Protein 90 (Hsp90) as a Molecular Target for Therapy of Gastrointestinal Cancer. Anticancer Res. 2009, 29, 2031. [Google Scholar]
  104. Mahalingam, D.; Swords, R.; Carew, J.S.; Nawrocki, S.T.; Bhalla, K.; Giles, F.J. Targeting HSP90 for cancer therapy. Br. J. Cancer 2009, 100, 1523–1529. [Google Scholar] [CrossRef] [Green Version]
  105. Neckers, L. Heat shock protein 90: The cancer chaperone. J. Biosci. 2007, 32, 517–530. [Google Scholar] [CrossRef]
  106. Li, Y.; Zhang, T.; Schwartz, S.J.; Sun, D. New developments in Hsp90 inhibitors as anti-cancer therapeutics: Mechanisms, clinical perspective and more potential. Drug Resist. Updat. 2009, 12, 17–27. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Mahajan, P.; Kaushal, J. Role of Phytoremediation in Reducing Cadmium Toxicity in Soil and Water. J. Toxicol. 2018, 2018, 4864365. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Chellaiah, E.R. Cadmium (heavy metals) bioremediation by Pseudomonas aeruginosa: A minireview. Appl. Water Sci. 2018, 8, 154. [Google Scholar] [CrossRef] [Green Version]
  109. Kumar, A.; Subrahmanyam, G.; Mondal, R.; Cabral-Pinto, M.M.S.; Shabnam, A.A.; Jigyasu, D.K.; Malyan, S.K.; Fagodiya, R.K.; Khan, S.A.; Yu, Z.G. Bio-remediation approaches for alleviation of cadmium contamination in natural resources. Chemosphere 2021, 268, 128855. [Google Scholar] [CrossRef]
  110. Fakhar, A.; Gul, B.; Gurmani, A.R.; Khan, S.M.; Ali, S.; Sultan, T.; Chaudhary, H.J.; Rafique, M.; Rizwan, M. Heavy metal remediation and resistance mechanism of Aeromonas, Bacillus, and Pseudomonas: A review. Crit. Rev. Environ. Sci. Technol. 2022, 52, 1868–1914. [Google Scholar] [CrossRef]
  111. Chasapis, C.T.; Peana, M.; Bekiari, V. Structural Identification of Metalloproteomes in Marine Diatoms, an Efficient Algae Model in Toxic Metals Bioremediation. Molecules 2022, 27. [Google Scholar] [CrossRef]
  112. Shah, V.; Daverey, A. Phytoremediation: A multidisciplinary approach to clean up heavy metal contaminated soil. Environ. Technol. Innov. 2020, 18, 100774. [Google Scholar] [CrossRef]
  113. Raza, A.; Habib, M.; Kakavand, S.N.; Zahid, Z.; Zahra, N.; Sharif, R.; Hasanuzzaman, M. Phytoremediation of Cadmium: Physiological, Biochemical, and Molecular Mechanisms. Biology (Basel) 2020, 9. [Google Scholar] [CrossRef]
  114. Hidalgo, J.; Aschner, M.; Zatta, P.; Vasak, M. Roles of the metallothionein family of proteins in the central nervous system. Brain Res. Bull. 2001, 55, 133–145. [Google Scholar] [CrossRef] [PubMed]
  115. Bjørklund, G.; Shanaida, M.; Lysiuk, R.; Antonyak, H.; Klishch, I.; Shanaida, V.; Peana, M. Selenium: An Antioxidant with a Critical Role in Anti-Aging. Molecules 2022, 27. [Google Scholar] [CrossRef] [PubMed]
  116. Cardoso, B.R.; Roberts, B.R.; Bush, A.I.; Hare, D.J. Selenium, selenoproteins and neurodegenerative diseases. Metallomics 2015, 7, 1213–1228. [Google Scholar] [CrossRef] [Green Version]
  117. Liu, M.C.; Xu, Y.; Chen, Y.M.; Li, J.; Zhao, F.; Zheng, G.; Jing, J.F.; Ke, T.; Chen, J.Y.; Luo, W.J. The effect of sodium selenite on lead induced cognitive dysfunction. Neurotoxicology 2013, 36, 82–88. [Google Scholar] [CrossRef]
  118. Sitek, A.; Kozlowska, L. The role of well-known antioxidant vitamins in the prevention of cadmium-induced toxicity. Int. J. Occup. Med. Environ. Health 2022, 35, 367–392. [Google Scholar] [CrossRef] [PubMed]
  119. Halttunen, T.; Collado, M.C.; El-Nezami, H.; Meriluoto, J.; Salminen, S. Combining strains of lactic acid bacteria may reduce their toxin and heavy metal removal efficiency from aqueous solution. Lett. Appl. Microbiol. 2008, 46, 160–165. [Google Scholar] [CrossRef]
  120. Forsyth, C.B.; Farhadi, A.; Jakate, S.M.; Tang, Y.; Shaikh, M.; Keshavarzian, A. Lactobacillus GG treatment ameliorates alcohol-induced intestinal oxidative stress, gut leakiness, and liver injury in a rat model of alcoholic steatohepatitis. Alcohol 2009, 43, 163–172. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  121. Zhai, Q.; Wang, G.; Zhao, J.; Liu, X.; Tian, F.; Zhang, H.; Chen, W. Protective effects of Lactobacillus plantarum CCFM8610 against acute cadmium toxicity in mice. Appl. Environ. Microbiol. 2013, 79, 1508–1515. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  122. Zhu, J.; Yu, L.; Shen, X.; Tian, F.; Zhao, J.; Zhang, H.; Chen, W.; Zhai, Q. Protective Effects of Lactobacillus plantarum CCFM8610 against Acute Toxicity Caused by Different Food-Derived Forms of Cadmium in Mice. Int. J. Mol. Sci. 2021, 22. [Google Scholar] [CrossRef]
  123. Zhai, Q.; Wang, G.; Zhao, J.; Liu, X.; Narbad, A.; Chen, Y.Q.; Zhang, H.; Tian, F.; Chen, W. Protective effects of Lactobacillus plantarum CCFM8610 against chronic cadmium toxicity in mice indicate routes of protection besides intestinal sequestration. Appl. Environ. Microbiol. 2014, 80, 4063–4071. [Google Scholar] [CrossRef] [Green Version]
  124. Deng, X.; Xia, Y.; Hu, W.; Zhang, H.; Shen, Z. Cadmium-induced oxidative damage and protective effects of N-acetyl-L-cysteine against cadmium toxicity in Solanum nigrum L. J. Hazard. Mater 2010, 180, 722–729. [Google Scholar] [CrossRef] [PubMed]
  125. Gil, H.W.; Kang, E.J.; Lee, K.H.; Yang, J.O.; Lee, E.Y.; Hong, S.Y. Effect of glutathione on the cadmium chelation of EDTA in a patient with cadmium intoxication. Hum. Exp. Toxicol. 2011, 30, 79–83. [Google Scholar] [CrossRef]
  126. Wu, X.; Su, S.; Zhai, R.; Chen, K.; Jin, T.; Huang, B.; Zhou, Y.; Ge, X.; Wei, G.; Liao, R. Lack of reversal effect of EDTA treatment on cadmium induced renal dysfunction: A fourteen-year follow-up. Biometals 2004, 17, 435–441. [Google Scholar] [CrossRef] [PubMed]
  127. Gonick, H.C. Nephrotoxicity of cadmium & lead. Indian J. Med. Res. 2008, 128, 335–352. [Google Scholar]
  128. Jalilehvand, F.; Leung, B.O.; Mah, V. Cadmium(II) complex formation with cysteine and penicillamine. Inorg. Chem. 2009, 48, 5758–5771. [Google Scholar] [CrossRef] [Green Version]
  129. Patrick, L. Toxic metals and antioxidants: Part II. The role of antioxidants in arsenic and cadmium toxicity. Altern. Med. Rev. A J. Clin. Ther. 2003, 8, 106–128. [Google Scholar]
  130. Flora, S.J.; Pachauri, V. Chelation in metal intoxication. Int. J. Environ. Res. Public Health 2010, 7, 2745–2788. [Google Scholar] [CrossRef] [Green Version]
  131. Rafati Rahimzadeh, M.; Kazemi, S.; Moghadamnia, A.A. Cadmium toxicity and treatment: An update. Caspian J. Intern. Med. 2017, 8, 135–145. [Google Scholar] [CrossRef]
  132. Routzomani, A.; Lada, Z.G.; Angelidou, V.; C, P.R.; Psycharis, V.; Konidaris, K.F.; Chasapis, C.T.; Perlepes, S.P. Confirming the Molecular Basis of the Solvent Extraction of Cadmium(II) Using 2-Pyridyl Oximes through a Synthetic Inorganic Chemistry Approach and a Proposal for More Efficient Extractants. Molecules 2022, 27. [Google Scholar] [CrossRef]
  133. Mazarakioti, E.C.; Beobide, A.S.; Angelidou, V.; Efthymiou, C.G.; Terzis, A.; Psycharis, V.; Voyiatzis, G.A.; Perlepes, S.P. Modeling the Solvent Extraction of Cadmium(II) from Aqueous Chloride Solutions by 2-pyridyl Ketoximes: A Coordination Chemistry Approach. Molecules 2019, 24. [Google Scholar] [CrossRef]
Figure 1. Examples of Cd2+ coordination environment in biological systems: (a) regular tetrahedral Cd2+ complexed with Hah1 metallochaperone protein (Protein Data Bank PDB 1FE0, [21]); (b) regular octahedral Cd2+ complexed with cytochrome c oxidase (PDB 2EIK, [22]); (c) a cluster of Cd2+ complexed with metallothionein-1 (PDB 1DFT, [23]).
Figure 1. Examples of Cd2+ coordination environment in biological systems: (a) regular tetrahedral Cd2+ complexed with Hah1 metallochaperone protein (Protein Data Bank PDB 1FE0, [21]); (b) regular octahedral Cd2+ complexed with cytochrome c oxidase (PDB 2EIK, [22]); (c) a cluster of Cd2+ complexed with metallothionein-1 (PDB 1DFT, [23]).
Biomolecules 13 00036 g001
Figure 2. Percentage distribution of geometries for mononuclear Cd2+ complexes in proteins, according to the Research Collaboratory for Structural Bioinformatics Protein Data Bank (RCSB PDB).
Figure 2. Percentage distribution of geometries for mononuclear Cd2+ complexes in proteins, according to the Research Collaboratory for Structural Bioinformatics Protein Data Bank (RCSB PDB).
Biomolecules 13 00036 g002
Figure 3. Percentage distribution of metal ions substituted by Cd2+, according to the Research Collaboratory for Structural Bioinformatics Protein Data Bank (RCSB PDB).
Figure 3. Percentage distribution of metal ions substituted by Cd2+, according to the Research Collaboratory for Structural Bioinformatics Protein Data Bank (RCSB PDB).
Biomolecules 13 00036 g003
Figure 4. Outcomes of cadmium exposure.
Figure 4. Outcomes of cadmium exposure.
Biomolecules 13 00036 g004
Figure 5. The 12 cancer types associated with specific gene-encoding Cd-metalloproteomes.
Figure 5. The 12 cancer types associated with specific gene-encoding Cd-metalloproteomes.
Biomolecules 13 00036 g005
Table 1. Names, UniProt IDs, abundances, and tissue-specific expressions of Cd-binding proteins, annotated as genetic markers in 12 cancer types. Abundances and tissue-specific expressions were mined according to the protein abundance database PaxDB (https://pax-db.org (accessed on 21 October 2022)) and human proteome atlas database (https://www.proteinatlas.org (accessed on 21 October 2022)).
Table 1. Names, UniProt IDs, abundances, and tissue-specific expressions of Cd-binding proteins, annotated as genetic markers in 12 cancer types. Abundances and tissue-specific expressions were mined according to the protein abundance database PaxDB (https://pax-db.org (accessed on 21 October 2022)) and human proteome atlas database (https://www.proteinatlas.org (accessed on 21 October 2022)).
A/AGeneUniProt IDAbundance (ppm)Protein NameHigh Expression Level
1GALNT10Q86SR1161Polypeptide N-acetylgalactosaminyltransferase 10Lung, gall bladder, kidney
2NOTCH4Q9946647Neurogenic locus notch homolog protein 4Adipose, lung
3GZMBP10144288Granzyme BDendritic cells
4BLKP5145136Tyrosine-protein kinase BlkSpleen, bone marrow
5AXLP30530681Tyrosine-protein kinase receptor UFOTestis, skeletal muscle
6TGFBR2P37173182TGF-beta receptor type-2Adipose, breast
7IL2RGP31785105Cytokine receptor common subunit gammaSpleen, tonsil
8LCKP06239941Tyrosine-protein kinase LckThymus, T-cells
9ESR1P0337235Estrogen receptorEndometrium, cervix, uterine
10PRKACAP17612697cAMP-dependent protein kinase catalytic subunit alphaCerebral cortex, testis
11SHCP29353200SHC-transforming protein 1Cerebellum, thyroid gland
12HSP90AB1P0823816962Heat shock protein HSP 90-betaCerebellum, adrenal gland
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Peana, M.; Pelucelli, A.; Chasapis, C.T.; Perlepes, S.P.; Bekiari, V.; Medici, S.; Zoroddu, M.A. Biological Effects of Human Exposure to Environmental Cadmium. Biomolecules 2023, 13, 36. https://doi.org/10.3390/biom13010036

AMA Style

Peana M, Pelucelli A, Chasapis CT, Perlepes SP, Bekiari V, Medici S, Zoroddu MA. Biological Effects of Human Exposure to Environmental Cadmium. Biomolecules. 2023; 13(1):36. https://doi.org/10.3390/biom13010036

Chicago/Turabian Style

Peana, Massimiliano, Alessio Pelucelli, Christos T. Chasapis, Spyros P. Perlepes, Vlasoula Bekiari, Serenella Medici, and Maria Antonietta Zoroddu. 2023. "Biological Effects of Human Exposure to Environmental Cadmium" Biomolecules 13, no. 1: 36. https://doi.org/10.3390/biom13010036

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop