Skip to main content

Toxic metals in the regulation of epithelial–mesenchymal plasticity: demons or angels?

Abstract

Epithelial cells can trans-differentiate into motile mesenchymal cells through a dynamic process known as epithelial–mesenchymal transition (EMT). EMT is crucial in embryonic development and wound healing but also contributes to human diseases such as organ fibrosis and cancer progression. Heavy metals are environmental pollutants that can affect human health in various ways, including causing cancers. The cytotoxicity and carcinogenicity of heavy metals are complex, and studies have demonstrated that some of these metals can affect the progress of EMT. Here, we focus on reviewing the roles of six environmentally common toxic metals concerning EMT: arsenic (AS), cadmium (Cd), cobalt (Co), chromium (Cr), nickel (Ni), and copper (Cu). Noteworthily, the effects of these elements on EMT may vary according to the form, dose, and exposure time; the dual role of heavy metals (e.g., AS, Cd, and Cu) on EMT is also observed, in which, sometimes they can promote while sometimes inhibit the EMT process. Given the vast number of toxicologically relevant metals that exist in nature, we believe a comprehensive understanding of their effects on EMT is required to dictate in what circumstances these metals act more likely as demons or angels.

Background

It is becoming very common for humans to expose to toxicologically relevant metals due to the diverse applications of metals in agriculture, medicine, household, technology, and industry [1]. Exposure to toxic metals has now permeated into all aspects of our lives, and not just from toxic waste sites or sporadic poisoning events. Thus, it is more important than ever before to pay attention to the potential adverse effects of metals on the environment as well as human health. In fact, the harmful effects of certain metals have only been noticed over the past few decades following the increase of human exposure worldwide owing to industrialization [2]. One of the reasons for slow recognition of metals’ perniciousness to human health is that the toxic effects are usually not instant and can take ages to accumulate. Therefore, even after knowing the potentially toxic effects, people are still willing to take the risks of using certain toxicologically relevant metals and alloys for the need of manufacture and ease of life. For example, dental “silver” amalgam fillings that contain about 50% elemental mercury are traditionally and are still commonly used in oral treatments despite having debatable safety issues [3].

Over the years, a great number of metallic elements have successively been proven to be carcinogenic based on epidemiological, clinical, in vitro, and in vivo studies [3]. All these metals have been classified as human carcinogens (either known or probable) by reputable organizations such as the International Agency for Research on Cancer (IARC) and the United States Environmental Protection Agency (USEPA) [1]. Epithelial–mesenchymal transition (EMT), a dynamic process where epithelial cells acquire mesenchymal features, is involved in developmental and morphogenetic processes but also contributes to human diseases such as organ fibrosis and cancer progression, especially metastasis [4, 5]. Recently, heavy metals such as arsenic (AS), cadmium (Cd), cobalt (Co), chromium (Cr), nickel (Ni), and copper (Cu) have been shown to play a role in EMT, and this review aims to provide a more holistic view of the effects of these metals in the progress of EMT.

What is EMT? Complex regulatory networks of EMT

Epithelial cells can trans-differentiate into motile mesenchymal cells through EMT, a morphogenetic process associated with wound healing, embryonic development, tissue formation, stem cell behavior, and cancer metastasis [6]. Conversely, a reversed process of EMT, known as mesenchymal–epithelial transition (MET), occurs when mesenchymal cells loss their migratory freedom and shift toward the epithelial state [6]. The concept of EMT was first described by Professor Elizabeth D. Hay in the early 1980s, as she observed the phenotypic changes of epithelial to mesenchymal state in the primitive streak of chick embryos [7]. Since then, EMT has attracted considerable attention in the field of cell biology as well as cancer research—it is now more than 20 years since EMT was first shown to be strongly associated with cancer progression [8, 9].

EMT is a complicated process involving more than hundreds of protein-coding and non-coding genes [10]. Some of these genes are selected and widely used as EMT markers, and these “classical” epithelial and mesenchymal markers are summarized in Fig. 1. In general, the simplest description of EMT from a molecular aspect is the loss of E-cadherin (CDH1; a cell–cell adhesion protein) and the gain of vimentin (VIM; a type-III intermediate filament protein) [6]. However, since EMT/MET occurs in a gradual manner, several intermediate states between the transition have been recently suggested, and these states can be classified as the partial-, incomplete-, and hybrid-EMT states in addition to the fully epithelial or mesenchymal state (reviewed in [6]). In many tumors, diverse EMT states of cancer cells have been observed, and these cells are associated with different metastatic potentials [11, 12].

Fig. 1
figure 1

An overview of the EMT process and classic gene markers. The EMT is a dynamic and reversible process modulated by epithelial and mesenchymal marker expression: Some of the typical epithelial markers include β-catenin (CTNNB1) and E-cadherin (CDH1), whereas mesenchymal markers include N-cadherin (CDH2), SNAI1/2 (SNAI1/2), and vimentin (VIM). Studies have shown that toxicologically relevant metals such as AS, Cd, Co, Cr, Ni, and Cu can promote the progress of EMT, and three of these metals (AS, Cd, and Cu) may inhibit EMT

The activation of the EMT transcription program is induced by signaling pathways mediated by transforming growth factor β (TGF-β), bone morphogenetic protein (BMP), Wnt/β-catenin, Notch, Hedgehog, and receptor tyrosine kinases [13, 14]. These pathways are usually triggered by various stimuli in the local microenvironment, such as growth factors, cytokines, hypoxia, and contact with the surrounding extracellular matrix [14]. Among the EMT-inducing pathways, the TGF-β signaling pathway is the most well-characterized, and it is usually activated by TGF-β superfamily ligands, including three TGF-β isoforms (β1, β2, and β3) and six BMP isoforms (BMP2 to BMP7) [14]. The Wnt/β-catenin signaling pathway is also considered to be a key signaling pathway driving EMT, which is mainly mediated by the Frizzled and low-density lipoprotein receptor-related protein (LRP) receptors. These Wnt receptor proteins can stabilize cytoplasmic β-catenin by preventing β-catenin from being phosphorylated by GSK-3β and degraded by proteasomes, resulting in the translocation of stabilized β-catenin to the nucleus and therefore activation of EMT [15]. Furthermore, in addition to the classical Wnt/β-catenin signaling pathway, Dissanayake et al. showed a noncanonical Wnt signaling pathway (Wnt5A/PKC signaling) that could also induce EMT [16].

The effect of heavy metals in EMT–demons or angels?

Most researchers would have agreed that almost all toxicologically relevant metals are evil because these toxic metals usually promote or exacerbate the process of EMT upon long-term and short-term exposures. However, analysis of recent experimental results shows that certain toxic metals can also inhibit EMT, depending on the dose and exposure time, and therefore these metals are portrayed as a combination of angels and demons. Take AS for example, although studies have shown that AS (III) exposure promotes intestinal tumor cell proliferation and invasion associated with EMT, exposure to low levels of AS (III) may also disrupt normal wound healing and angiogenesis processes of metastatic cancer cells [17, 18]. The current knowledge regarding the EMT promoting or inhibitory effects of AS, Cd, Co, Cr, Ni, and Cu is summarized in Table 1.

Table 1 An overview of the effect of the six toxicologically relevant metals in EMT

Arsenic

AS is widely distributed in the environment due to its natural existence and industrial and medical applications. The major inorganic forms of AS are the pentavalent arsenate and trivalent arsenite. The organic forms include the methylated metabolites monomethylarsonous acid (MMA), dimethylarsinic acid, and trimethylarsine oxide. Although there is AS pollution in the air, soil, and other sources, people are mainly exposed to unsafe levels of AS through contaminated drinking water [19]. Exposure to AS can cause serious health consequences, such as cardiovascular disease, conjunctival congestion, diabetes mellitus, weakness, neurological deficits, hypertension, cancer, and other chronic diseases [20].

Evidence is clear that AS can induce EMT in normal epithelial cells of various organs, even at low concentrations. For example, studies have indicated that chronic exposure to low levels of AS (1.0–2.5 μM of NaAsO2) resulted in human bronchial epithelial (HBE) cells to acquire stem cell-like properties and malignant transformation, in which, these changes were shown to be associated with the deletion of miR-200 family members and upregulation of miR-21 that induced EMT [21,22,23]. Furthermore, several experiments have demonstrated that HBE or human epidermal keratinocyte (HaCaT) cells chronically exposed to NaAsO2 resulted in increased IL-6 and miR-21, decreased PTEN, and activation of STAT3 and AKT signaling pathways [24,25,26,27,28]. Further mechanistic studies revealed that EMT activated β-catenin in AS-transformed HBE cells, which upregulated the level of angiogenic-stimulating growth factor VEGF and promoting angiogenesis [29].

In human bronchial epithelial BEAS-2B cells, chronic exposure to sub-lethal doses of NaAsO2 resulted in the inhibition of miR-100 expression, activation of autophagy, and induction of EMT via the MEK/ERK1/2 signaling pathway [30, 31]. It has been shown that miR-191 is a highly conserved oncogenic miRNA [32]. In human liver epithelial L-02 cells, NaAsO2 exposure increased the level of HIF-2α-mediated miR-191, and thus promoting EMT and cancer stem cell-like phenotypes [33]. In another study, treatment with various concentrations (0–8 μM) of NaAsO2 for 12 h or with 2 μM NaAsO2 for selected periods (0–24 h) in L-02 cells indicated that AS could increase the number of autophagosomes by blocking autophagic flux, leading to the accumulation of SQSTM1/p62 and upregulation of mesenchymal protein SNAI1 [34]. Similarly, treatment of human lung peripheral epithelial cells (HPL-1D) and human immortalized uroepithelial cells (SV-HUC-1) with low levels of NaAsO2 can induce EMT [35, 36]. By exposing NaAsO2 to renal cortex/proximal tubule (HK-2) epithelial cells for 72 h (acute), 3 months (long-term), and 6 months (chronic), Chang and Singh found that HK-2 cells could undergo neoplastic transformation through the acquisition of EMT when chronically exposed to a relatively lower concentration of AS [37]. Furthermore, they also found that long-term AS exposure could cause HK-2 cells to acquire DNA methylation-mediated fibrotic phenotypes and treatment of DNA methylation inhibitor 5-Aza-2'-dC could reverse the EMT properties [38].

In additional to normal epithelial cells, AS-induced EMT can also be observed in other cell types, including cancer cells. For instance, in colorectal cancer cell lines Caco2 and HCT116, exposure to NaAsO2 could induce SEPT9 promoter hypomethylation, which further initiated EMT [39]. In another study, long-term treatment (6 months) of immortalized human keratinocytes (NHEK/SVTERT3-5) cells with AS trioxide (ATO) could induce EMT, impair differentiation of organotypic skin models, and mimic aspects of human skin derangements (e.g., Bowen’s disease) [40]. Also, EMT and global methylation changes were observed in human cervical cancer HeLa cells when treated with 0.5 µM NaAsO2 for about 45 days [41].

Although the above data have shown that AS can promote EMT and enhance tumor malignancy, some contradictory results indicate that AS can suppress EMT and may be a potent anticancer agent. Several studies have indicated that low levels of AS (III) may inhibit wound healing and angiogenesis of metastatic cancer cells [17]. In gastric cancer cell lines (AGS and MGC803), As4S4 treatment could upregulate the expression of miR-4665-3p, which in turn downregulated the expression of oncoprotein GSE1 and resulted in the reverse of EMT in these cell lines [42]. In another study, EMT in AGS cells was inhibited by ATO treatment (5 or 10 μM, 48 h) via the induction of SHP-1 and attenuation of p-JAK2/p-STAT3 [43]. ATO was also shown to suppress EMT, tumor progression, and metastasis in Buffalo rat hepatoma cell line Mca-Rh7777 by inhibiting TWIST activation [44]. Furthermore, ATO could weaken the invasiveness of chondrosarcoma cells and reverse the cells to more epithelial states by increasing the expression of miR-125b, a process associated with the demethylation of DNA [45]. In hepatocellular carcinoma cell lines, ATO treatment was able to inhibit EMT by suppressing the expression of PKM2 via the induction of anticancer lncRNA MEG3 [46]. Overall, based on these results, we can observe that exposure to AS could promote angiogenesis and EMT, leading to malignant transformation of cells as well as enhancing the migration and invasion of tumor cells; on the other hand, short-term AS exposure, especially ATO, may inhibit tumor progression and metastasis and therefore have clinical implications. However, caution should be taken when applying AS for clinical used since the EMT inhibition properties of AS could also cause other non-cancer disorders, particularly cardiovascular diseases: it was indicated that MMA (III) could inhibit EMT of epicardial cells that result in AS-associated cardiovascular disorders [47]. Also, Allison et al. showed that NaAsO2 exposure could disrupt TGF-β2 signals and Smad activation, leading to the blockage of developmental EMT gene programming in murine coronary progenitor cells, but AS toxicity had no significant effect on smooth muscle differentiation [48].

Cadmium

Cd is a toxic heavy metal with considerable effect on the environment and human health. As a naturally occurring element, the presence of Cd in the environment has been substantially magnified by industrialization and human activities. Humans are mainly exposed to Cd through a number of sources, including consumption of Cd-contaminated food, working in Cd-contaminated workplaces, and smoking of cigarettes [49]. Findings from repeated studies of occupational Cd exposure and lung cancer have concluded that Cd is a human carcinogen according to the IARC and the USEPA. Some studies have also determined that Cd exposure is associated with cancers of the prostate, kidney, liver, hematopoietic system, and stomach [50].

Given the fact that Cd exposure (even through the gastrointestinal system) is strongly correlated to lung cancer, many studies have used lung cell models to study Cd toxicity [51, 52]. In our previous studies, we found that BEAS-2B cells chronically exposed to CdCl2 exhibited EMT phenotype that ubiquitin carboxyl-terminal hydrolase isozyme L1, a newly identified EMT suppressor, was severely downregulated in the Cd-resistant BEAS-2B cell model [53, 54]. In another study by Tanwar et al., short-term Cd exposure (0, 2.5, 5, and 10 μM CdCl2 for 72 h) was able to decrease the level of miR-30 family genes and upregulated SNAI1 in human lung epithelial cells [55]. In addition to human lung cells, Cd could also induce EMT in tissues of other organs: it was shown that non-cancerous breast (MCF10A) and pancreas (hTERT-HPNE) epithelial cell lines underwent EMT after exposure to 2.5 µM CdCl2 for 40 weeks [56]. As a key regulator of EMT, SNAI1 was also found to be upregulated upon treatment with 1 μM or 3 μM CdCl2 for 4 weeks in both normal and cancer-derived breast epithelial cells [57]. Furthermore, Shan et al. showed that triple-negative human breast cancer cell line MDA-MB-231 treated with 1–3 μM of CdCl2 for 8 weeks resulted in the suppression of breast cancer prognostic marker ferroportin, increased intracellular iron concentration, promotion of cell proliferation and migration, and induction of EMT [58]. Similarly, in other cancer cell lines, Cd-associated EMT was observed. For example, in renal cancer Caki-1 cells, EMT was promoted by Cd treatment via upregulation of PGE2 through cAMP/PKA-COX2 signaling pathway [59]. In lung adenocarcinoma A549 cells, prolonged CdCl2 exposure induced EMT and malignant progression via the activation of Notch1, hypoxia-inducible factor-1α (HIF-1α), and IGF-1R/Akt/ERK/S6K1 signaling pathways [60]. The induction of EMT by Cd described above has also been supported by animal studies. In order to mimic long term and chronic Cd exposure, Chakraborty et al. used drinking water containing environmentally relevant Cd (100 mg/l of CdCl2) to feed mice for 12 weeks, and renal fibrosis was observed, along with activation of the Wnt pathway and increased expression of EMT-related genes in the kidney tissues [61].

Despite overwhelming evidence indicating that Cd promotes EMT, one study has suggested that “physiologically” relevant concentrations of CdCl2 (0.25 and 2.5 μM) can inhibit EMT in adult mammary stem cells via the inhibition of HIF-1α activity (important for human mammary stem cell proliferation and branching morphogenesis) and downregulation of EMT-associated genes (e.g., VIM, ZEB1, and TGFBI) [62].

Cobalt

As a rare element, the chemical properties of Co are highly similar to iron and Ni. Co can form stable salts and complex compounds, mainly as Co (II) oxide and Co (III) oxide [63]. Humans are often exposed to a wide variety of Co compounds due to their widespread occurrence in daily life, including occupational, environmental, dietary, and medical.

CoCl2 is a hypoxia-mimetic agent commonly used to simulate the typical hypoxic environment of cancer cells. Therefore, it is also often used to study the role of hypoxia in cancer development [64, 65]. Studies have indicated that CoCl2 can promote EMT by stabilizing HIF-1α (a key hypoxia marker) in various cancer cell lines. For example, stabilization of HIF-1α was observed in human pancreatic carcinoma (MiaPaCa2) and esophageal squamous cell carcinoma (TE-1 and EC-1) cell lines treated with CoCl2, and it was shown that the hypoxic environment in these cell lines promoted EMT via the activation of Notch1-STAT3 signaling pathway, downregulation of E-cadherin, and increased expression of N-cadherin and SNAI1 [66, 67]. In human hepatocellular carcinoma HepG2 cell line, HIF-1α increased the level of COX-2 protein and induced EMT process to cope with hypoxic environment, leading to increased invasiveness and metastasis of the cancer cells [68].

Similarly, human lung cancer cell lines (A549 and PC9) treated with 100 µM CoCl2 for 24–48 h exhibited EMT phenotypes such as increased invasion and migration, and these cell lines also showed increased expressions of Netrin-1 and vimentin, activated PI3K/AKT pathway, and downregulation of E-cadherin [69]. In breast cancer, Chu et al. showed that expressions of vimentin and matrix metalloproteinases (MMP2 and MMP9) were significantly increased due to hypoxia in ductal carcinoma (MDA-MB-231) and mammary tumor (MCF7) cell lines treated with CoCl2 [70]. The expression of CA IX, a novel prognostic marker protein for breast cancer, was also upregulated in these breast cancer cell lines and closely related to tumor cell migration and invasion [70]. In another study, it was indicated that long-term treatment of CoCl2 could increase the number of polyploid giant cancer cells, and these cells could asymmetrically divide into more aggressive daughter cells in breast cancer [71]. Additionally, Lester et al. showed that expression of urokinase-type plasminogen activator receptor (uPAR) was induced by hypoxia in the breast cancer MDA-MB-231 cells. The overexpression of uPAR activated uPAR-dependent cell signaling and promoted EMT in the cancer cells, and this process could be reversed by silencing the expression of uPAR or by blocking the uPAR-activated cell signaling factors [72]. Furthermore, Thongchot et al. indicated that the HIF-1α expression in and cell migration of cholangiocarcinoma cell lines (M139 and M214) with CoCl2-stimulated hypoxia conditions could be suppressed by treating the cells with chloroquine [73].

Besides cancer cell lines, evidence is also clear that CoCl2 can induce EMT in various types of normal human cells [74, 75]. For instance, Kong et al. showed that human LO2 hepatocytes treated with 100 μM for 24 or 72 h underwent EMT, and this process could be inhibited by curcumin treatment via TGF-β/Smad signaling interference [74]. In another study, human lens epithelium cell line (SRA01/04) treated with 150 μM CoCl2 for 48 h exhibited EMT properties such as decreased expression of E-cadherin, increased expressions of HIF-1α and Notch1, activation of SNAI1, and enhanced cell migration [75].

Chromium

Cr is an element naturally present in the earth's crust with several oxidation states, and the two more common ones are Cr (III) and Cr (VI) [76]. The oxidation state of Cr dictates the health hazard of Cr exposure: Cr (III) is non-toxic and good for nutrition and health whereas Cr (VI) is extremely toxic and has been classified as a group I carcinogen by the IARC and USEPA [3]. In general, humans are exposed to Cr (VI) through the ingestion of contaminated food/water and occupational inhalation [77, 78]—it is estimated that more than 300,000 workers are exposed to Cr (VI) and Cr-containing compounds in the workplace each year [79].

Exposure to Cr (VI) may induce health problems, including damage to the sperm and male reproductive system, anemia, and higher risk of cancers [1]. It was discovered that the Cr (VI) concentrations in the serum of prostate cancer patients were much higher than those of benign prostatic hyperplasia patients [80]. Further in vitro and in vivo studies revealed that exposure to low doses of Cr (VI) might affect prostate cancer progression by inducing EMT [80]. The EMT-inducing ability of Cr (VI) is not only observed in cancer cells but also in normal human cell lines. For instance, it was indicated that suppressed E-cadherin levels, increased vimentin levels, and EMT phenotypes (e.g., fibroblastoid morphology) were associated with acute and chronic K2Cr2O7 exposures in the BEAS-2B cells [81]. In another study, Li et al. found that K2Cr2O7 could increase the levels of mesenchymal protein and stem cell markers in renal epithelial cells [82].

Nickel

Ni, as the 2nd most abundant element in the Earth’s inner core, is widely distributed in the environment, air, water, and soil [3]. Ni is used in a broad variety of metallurgical processes and as a catalyst in the chemical and food industry [83]. Humans are exposed to Ni mainly through Ni-contaminated water and food, which can cause a variety of health hazards. Depending on the dose and duration of exposure, Ni can cause cardiovascular diseases, lung fibrosis, and cancer of the respiratory tract [84, 85].

The relationships between Ni exposure and lung tumorigenesis have been assessed by several researchers. Wu et al. showed that treatment of NiCl2 could induce fibronectin and promote TGF-β-induced EMT by decreasing the level of TAB2 via upregulation of miR-4417, in both normal (BEAS-2B) and cancerous (A549) human lung cell lines [86]. Similarly, in another study by Jose et al., EMT was induced in the BEAS-2B cells chronically exposed to 100 μM NiCl2 for 6 weeks, and the persistent gene expression changes in the Ni-treated cells were examined [87]. They found that upregulation of ZEB1 was required for Ni-induced EMT, and the expression of ZEB1 was persistently activated by Ni-induced epigenetic alterations (e.g., decreased H3K27me3 levels) but not by hypoxia [87].

Copper

As an essential metal, Cu plays key roles in many physiological functions, such as oxidation resistance, energy metabolism, neuronal function, and tissue integrity [88]. However, there is also evidence indicating that excessive Cu will induce angiogenesis because Cu can directly or indirectly regulate numerous angiogenesis-related factors [89]. In addition, recent studies have suggested that Cu could be carcinogenic, and Cu exposure may be associated with breast, lung, brain, colon, and prostate cancers [90].

In general, exposure to CuCl2 can lead to transactivation of EMT marker genes by increasing the activity of HIF-1α [91, 92]. Guo et al. showed that CuSO4 could induce EMT via activation of TGF-β1/Smad and MAPKs pathways in the lung of CuSO4-treated mice, resulting in pulmonary fibrosis [93]. On the other hand, Li et al. showed that the removal of Cu through the silencing of Ctr1 (a transmembrane protein responsible for cellular Cu uptake) could inhibit CoCl2-induced EMT via HIF-1α de-stabilization, along with SNAI1 and TWIST downregulation [90]. Therefore, it has been suggested that Cu chelators have the potential to be established as anticancer drugs worthy of clinical consideration.

Although most studies have indicated that exposure to high doses of Cu is a cancer risk, some studies have suggested the opposite effects of Cu in cancer. Specifically, it was revealed that disulfiram (DSF), an aldehyde dehydrogenase inhibitor with anticancer activity [94], displayed improved anti-angiogenic activity in a Cu-dependent manner [95]. Further study on this matter showed that DSF combined with Cu could suppress hepatic carcinoma metastasis and EMT by repressing NF-κB and TGF-β1 signaling pathways [96]. Overall, the above results indicate that the effects of Cu seem to be promiscuous as Cu exhibits carcinogenic and antitumor properties.

Risk and opportunity

The numbers and levels of heavy metals in the living environment have risen dramatically over the years owing to a series of human activities, including technological advancement, urbanization, rapid industrialization, and unsafe agricultural practices [97]. As a result, exposure to heavy metals has become a serious global health problem as these elements can accumulate in the body and cause various human diseases, including cancers [98,99,100]. Currently, humans are mainly exposed to toxicologically relevant metals through several sources, including diet, polluted air, occupational inhalation, and cigarette smoking.

It is learned that in recent years, the occurrences of heavy metal-contaminated food (e.g., Cd rice) due to water/land pollution have increased significantly on a global scale. Thus, in order to reduce the risk of toxic metals, more systematic research on food, workplace, and environmental safety is required. Meanwhile, there are several methods that may help us reduce our exposure to heavy metals, and the most critical one is that we control and, if possible, eliminate the pollution sources. From an agricultural and industrial aspect, it is important that we rationalize the usage of fertilizers and pesticides and strictly prohibit the discharge of industrial wastes. From a government aspect, relevant departments shall formulate and improve the workplace and environmental regulations and standards, strengthen supervision, and raise public awareness of environmental protection. From a personal daily life aspect, we can use a drinking water filtration system to filter heavy metals, take precautions at work, and avoid going to or living in highly polluted areas.

The cytotoxicity and carcinogenicity of heavy metals are complex, and recent studies have demonstrated that some of these metals can induce EMT in both normal and cancerous cells, leading to increased cancer risk (as shown in Fig. 2). However, as elaborated in this review, the dual role of certain heavy metals in cancer has also been observed—these heavy metals (e.g., AS, Cd, and Cu) exhibit anticancer properties, suggesting the opportunity for them to be applied in cancer therapy.

Fig. 2
figure 2

Health risks associated with the six heavy metals reviewed in this article. Chronic exposure to these metals could promote EMT and the development of cancers in the lung, breast, liver, kidney, bladder, and prostate

Conclusions

In recent years, knowledge in the field of EMT has continued to expand. Even though the contribution of metals and nutrients in human cancer has been well recognized for many years, it is now becoming clear that certain metals are able to impact the process of EMT. In this review, we provide a deeper and more comprehensive picture of the effects of six toxicologically relevant metals (AS, Cd, Co, Cr, Ni, and Cu) in EMT. We also show that most of the time, heavy metals act more likely as demons by promoting EMT, while sometimes they could act more likely as angels by inhibiting EMT (summarized in Table 2). Overall, the toxicity of heavy metals is expected to be decided by the dose, route, and chemical species of exposure, together with the nutritional status, gender, age, and genetics of exposed individuals. In conclusion, this review reminds the public that we should avoid or reduce the chance of exposure to toxic heavy metals in our daily lives as exposure to these metals may promote EMT and induce malignant transformation of cells.

Table 2 Toxicologically relevant metals-induced EMT marker alterations

Availability of data and materials

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

Abbreviations

AS:

Arsenic

ATO:

Arsenic trioxide

BMP:

Bone morphogenetic protein

Cd:

Cadmium

Co:

Cobalt

Cr:

Chromium

Cu:

Copper

DSF:

Disulfiram

EMT:

Epithelial–mesenchymal transition

HBE:

Human bronchial epithelial

HIF-1α:

Hypoxia-inducible factor-1α

MMA:

Monomethylarsonous acid

TGF-β:

Transforming growth factor β

uPAR:

Urokinase-type plasminogen activator receptor

Ni:

Nickel

References

  1. Tchounwou PB, Yedjou CG, Patlolla AK, Sutton DJ. Heavy metal toxicity and the environment. Exp Suppl. 2012;101:133–64.

    PubMed  Google Scholar 

  2. Rehman K, Fatima F, Waheed I, Akash MSH. Prevalence of exposure of heavy metals and their impact on health consequences. J Cell Biochem. 2018;119:157–84.

    Article  CAS  PubMed  Google Scholar 

  3. Chen QY, DesMarais T, Costa M. Metals and mechanisms of carcinogenesis. Annu Rev Pharmacol Toxicol. 2019;59:537–54.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Nakaya Y, Sheng G. EMT in developmental morphogenesis. Cancer Lett. 2013;341:9–15.

    Article  CAS  PubMed  Google Scholar 

  5. Thiery JP. Epithelial-mesenchymal transitions in development and pathologies. Curr Opin Cell Biol. 2003;15:740–6.

    Article  CAS  PubMed  Google Scholar 

  6. Nieto MA, Huang RY, Jackson RA, Thiery JP. EMT:2016. Cell. 2016;166:21–45.

    Article  CAS  PubMed  Google Scholar 

  7. Hay ED. An overview of epithelial mesenchymal transformation. Acta Anat (Basel). 1995;154:8–20.

    Article  CAS  Google Scholar 

  8. Savagner P, Yamada KM, Thiery JP. The zinc-finger protein slug causes desmosome dissociation, an initial and necessary step for growth factor-induced epithelial-mesenchymal transition. J Cell Biol. 1997;137:1403–19.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Thiery JP. Epithelial–mesenchymal transitions in tumour progression. Nat Rev Cancer. 2002;2:442–54.

    Article  CAS  PubMed  Google Scholar 

  10. Vasaikar SV, Deshmukh AP, den Hollander P, Addanki S, Kuburich NA, Kudaravalli S, Joseph R, Chang JT, Soundararajan R, Mani SA. EMTome: a resource for pan-cancer analysis of epithelial-mesenchymal transition genes and signatures. Br J Cancer. 2021;124:259–69.

    Article  CAS  PubMed  Google Scholar 

  11. Pastushenko I, Brisebarre A, Sifrim A, Fioramonti M, Revenco T, Boumahdi S, Van Keymeulen A, Brown D, Moers V, Lemaire S, et al. Identification of the tumour transition states occurring during EMT. Nature. 2018;556:463–8.

    Article  CAS  PubMed  Google Scholar 

  12. Thiery JP, Sleeman JP. Complex networks orchestrate epithelial-mesenchymal transitions. Nat Rev Mol Cell Biol. 2006;7:131–42.

    Article  CAS  PubMed  Google Scholar 

  13. Dongre A, Weinberg RA. New insights into the mechanisms of epithelial-mesenchymal transition and implications for cancer. Nat Rev Mol Cell Biol. 2019;20:69–84.

    Article  CAS  PubMed  Google Scholar 

  14. Gonzalez DM, Medici D. Signaling mechanisms of the epithelial-mesenchymal transition. Sci Signal. 2014;7:re8. https://doi.org/10.1126/scisignal.2005189.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Savagner P. Leaving the neighborhood: molecular mechanisms involved during epithelial-mesenchymal transition. Bioessays. 2001;23:912–23.

    Article  CAS  PubMed  Google Scholar 

  16. Dissanayake SK, Wade M, Johnson CE, O’Connell MP, Leotlela PD, French AD, Shah KV, Hewitt KJ, Rosenthal DT, Indig FE, et al. The Wnt5A/protein kinase C pathway mediates motility in melanoma cells via the inhibition of metastasis suppressors and initiation of an epithelial to mesenchymal transition. J Biol Chem. 2007;282:17259–71.

    Article  CAS  PubMed  Google Scholar 

  17. Lencinas A, Broka DM, Konieczka JH, Klewer SE, Antin PB, Camenisch TD, Runyan RB. Arsenic exposure perturbs epithelial-mesenchymal cell transition and gene expression in a collagen gel assay. Toxicol Sci. 2010;116:273–85.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Sun JL, Chen DL, Hu ZQ, Xu YZ, Fang HS, Wang XY, Kan L, Wang SY. Arsenite promotes intestinal tumor cell proliferation and invasion by stimulating epithelial-to-mesenchymal transition. Cancer Biol Ther. 2014;15:1312–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Mandal P. Molecular insight of arsenic-induced carcinogenesis and its prevention. Naunyn Schmiedebergs Arch Pharmacol. 2017;390:443–55.

    Article  CAS  PubMed  Google Scholar 

  20. Bjørklund G, Aaseth J, Chirumbolo S, Urbina MA, Uddin R. Effects of arsenic toxicity beyond epigenetic modifications. Environ Geochem Health. 2017;40:955–65.

    Article  PubMed  CAS  Google Scholar 

  21. Wang Z, Zhao Y, Smith E, Goodall GJ, Drew PA, Brabletz T, Yang C. Reversal and prevention of arsenic-induced human bronchial epithelial cell malignant transformation by microRNA-200b. Toxicol Sci. 2011;121:110–22.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Xu Y, Li Y, Pang Y, Ling M, Shen L, Yang X, Zhang J, Zhou J, Wang X, Liu Q. EMT and stem cell-like properties associated with HIF-2α are involved in arsenite-induced transformation of human bronchial epithelial cells. PLoS One. 2012;7:e37765.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Luo F, Ji J, Liu Y, Xu Y, Zheng G, Jing J, Wang B, Xu W, Shi L, Lu X, et al. MicroRNA-21, up-regulated by arsenite, directs the epithelial-mesenchymal transition and enhances the invasive potential of transformed human bronchial epithelial cells by targeting PDCD4. Toxicol Lett. 2015;232:301–9.

    Article  CAS  PubMed  Google Scholar 

  24. Jiang R, Li Y, Xu Y, Zhou Y, Pang Y, Shen L, Zhao Y, Zhang J, Zhou J, Wang X, et al. EMT and CSC-like properties mediated by the IKKβ/IκBα/RelA signal pathway via the transcriptional regulator, Snail, are involved in the arsenite-induced neoplastic transformation of human keratinocytes. Arch Toxicol. 2013;87:991–1000.

    Article  CAS  PubMed  Google Scholar 

  25. Lu X, Luo F, Liu Y, Zhang A, Li J, Wang B, Xu W, Shi L, Liu X, Lu L, et al. The IL-6/STAT3 pathway via miR-21 is involved in the neoplastic and metastatic properties of arsenite-transformed human keratinocytes. Toxicol Lett. 2015;237:191–9.

    Article  CAS  PubMed  Google Scholar 

  26. Xue J, Chen C, Luo F, Pan X, Xu H, Yang P, Sun Q, Liu X, Lu L, Yang Q, et al. CircLRP6 regulation of ZEB1 via miR-455 is involved in the epithelial-mesenchymal transition during arsenite-induced malignant transformation of human keratinocytes. Toxicol Sci. 2018;162:450–61.

    Article  CAS  PubMed  Google Scholar 

  27. Lu X, Liu Y, Luo F, Zhang A, Liu X, Lu L, Shi L, Li J, Xue J, Xu H, et al. MicroRNA-21 activation of Akt via PTEN is involved in the epithelial-mesenchymal transition and malignant transformation of human keratinocytes induced by arsenite. Toxicol Res (Camb). 2016;5:1140–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Luo F, Xu Y, Ling M, Zhao Y, Xu W, Liang X, Jiang R, Wang B, Bian Q, Liu Q. Arsenite evokes IL-6 secretion, autocrine regulation of STAT3 signaling, and miR-21 expression, processes involved in the EMT and malignant transformation of human bronchial epithelial cells. Toxicol Appl Pharmacol. 2013;273:27–34.

    Article  CAS  PubMed  Google Scholar 

  29. Wang Z, Humphries B, Xiao H, Jiang Y, Yang C. Epithelial to mesenchymal transition in arsenic-transformed cells promotes angiogenesis through activating β-catenin-vascular endothelial growth factor pathway. Toxicol Appl Pharmacol. 2013;271:20–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Wu L, Wang Z, Li X, He X, Han Y, Chen Y, Liu L, Fu L, Zhang T. Autophagy mediates bronchial cell malignant transformation induced by chronic arsenic exposure via MEK/ERK1/2 pathway. Toxicol Lett. 2020;332:155–63.

    Article  CAS  PubMed  Google Scholar 

  31. Yang J, Chen Z, Wang X, Xu M, Fang H, Li F, Liu Y, Jiang Y, Ding Y, Li J, et al. Inactivation of miR-100 combined with arsenic treatment enhances the malignant transformation of BEAS-2B cells via stimulating epithelial-mesenchymal transition. Cancer Biol Ther. 2017;18:965–73.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Nagpal N, Kulshreshtha R. miR-191: an emerging player in disease biology. Front Genet. 2014;5:99.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  33. Chen C, Yang Q, Wang D, Luo F, Liu X, Xue J, Yang P, Xu H, Lu J, Zhang A, et al. MicroRNA-191, regulated by HIF-2α, is involved in EMT and acquisition of a stem cell-like phenotype in arsenite-transformed human liver epithelial cells. Toxicol In Vitro. 2018;48:128–36.

    Article  CAS  PubMed  Google Scholar 

  34. Liu X, Ling M, Chen C, Luo F, Yang P, Wang D, Chen X, Xu H, Xue J, Yang Q, et al. Impaired autophagic flux and p62-mediated EMT are involved in arsenite-induced transformation of L-02 cells. Toxicol Appl Pharmacol. 2017;334:75–87.

    Article  CAS  PubMed  Google Scholar 

  35. Person RJ, Ngalame NN, Makia NL, Bell MW, Waalkes MP, Tokar EJ. Chronic inorganic arsenic exposure in vitro induces a cancer cell phenotype in human peripheral lung epithelial cells. Toxicol Appl Pharmacol. 2015;286:36–43.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Zhou Q, Jin P, Liu J, Li S, Liu W, Xi S. Arsenic-induced HER2 promotes proliferation, migration and angiogenesis of bladder epithelial cells via activation of multiple signaling pathways in vitro and in vivo. Sci Total Environ. 2021;753:141962.

    Article  CAS  PubMed  Google Scholar 

  37. Chang YW, Singh KP. Arsenic-induced neoplastic transformation involves epithelial-mesenchymal transition and activation of the β-catenin/c-Myc pathway in human kidney epithelial cells. Chem Res Toxicol. 2019;32:1299–309.

    Article  CAS  PubMed  Google Scholar 

  38. Chang YW, Singh KP. Arsenic induces fibrogenic changes in human kidney epithelial cells potentially through epigenetic alterations in DNA methylation. J Cell Physiol. 2019;234:4713–25.

    Article  CAS  PubMed  Google Scholar 

  39. Rafiei G, Shirkoohi R, Saffari M, Salehipour P, Modarressi MH. The impact of long-term exposure to low levels of inorganic Arsenic on the hypomethylation of SEPT9 promoter in epithelial-mesenchymal transformed colorectal cancer cell lines. Int J Mol Cell Med. 2019;8:130–8.

    CAS  PubMed  PubMed Central  Google Scholar 

  40. Weinmuellner R, Kryeziu K, Zbiral B, Tav K, Schoenhacker-Alte B, Groza D, Wimmer L, Schosserer M, Nagelreiter F, Rösinger S, et al. Long-term exposure of immortalized keratinocytes to arsenic induces EMT, impairs differentiation in organotypic skin models and mimics aspects of human skin derangements. Arch Toxicol. 2018;92:181–94.

    Article  CAS  PubMed  Google Scholar 

  41. Eckstein M, Rea M, Fondufe-Mittendorf YN. Transient and permanent changes in DNA methylation patterns in inorganic arsenic-mediated epithelial-to-mesenchymal transition. Toxicol Appl Pharmacol. 2017;331:6–17.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Zhang X, Tan Z, Kang T, Zhu C, Chen S. Arsenic sulfide induces miR-4665-3p to inhibit gastric cancer cell invasion and migration. Drug Des Devel Ther. 2019;13:3037–49.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Kim SH, Yoo HS, Joo MK, Kim T, Park JJ, Lee BJ, Chun HJ, Lee SW, Bak YT. Arsenic trioxide attenuates STAT-3 activity and epithelial-mesenchymal transition through induction of SHP-1 in gastric cancer cells. BMC Cancer. 2018;18:150.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  44. Wang GZ, Zhang W, Fang ZT, Zhang W, Yang MJ, Yang GW, Li S, Zhu L, Wang LL, Zhang WS, et al. Arsenic trioxide: marked suppression of tumor metastasis potential by inhibiting the transcription factor Twist in vivo and in vitro. J Cancer Res Clin Oncol. 2014;140:1125–36.

    Article  CAS  PubMed  Google Scholar 

  45. Bao X, Ren T, Huang Y, Wang S, Zhang F, Liu K, Zheng B, Guo W. Induction of the mesenchymal to epithelial transition by demethylation-activated microRNA-125b is involved in the anti-migration/invasion effects of arsenic trioxide on human chondrosarcoma. J Exp Clin Cancer Res. 2016;35:129.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  46. Fan Z, He J, Fu T, Zhang W, Yang G, Qu X, Liu R, Lv L, Wang J. Arsenic trioxide inhibits EMT in hepatocellular carcinoma by promoting lncRNA MEG3 via PKM2. Biochem Biophys Res Commun. 2019;513:834–40.

    Article  CAS  PubMed  Google Scholar 

  47. Huang T, Barnett JV, Camenisch TD. Cardiac epithelial-mesenchymal transition is blocked by monomethylarsonous acid (III). Toxicol Sci. 2014;142:225–38.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Allison P, Huang T, Broka D, Parker P, Barnett JV, Camenisch TD. Disruption of canonical TGFβ-signaling in murine coronary progenitor cells by low level arsenic. Toxicol Appl Pharmacol. 2013;272:147–53.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Tan HW, Liang ZL, Yao Y, Wu DD, Mo HY, Gu J, Chiu JF, Xu YM, Lau ATY. Lasting DNA damage and aberrant DNA repair gene expression profile are associated with post-chronic cadmium exposure in human bronchial epithelial cells. Cells. 2019;8:842.

    Article  CAS  PubMed Central  Google Scholar 

  50. Nawrot T, Plusquin M, Hogervorst J, Roels HA, Celis H, Thijs L, Vangronsveld J, Van Hecke E, Staessen JA. Environmental exposure to cadmium and risk of cancer: a prospective population-based study. Lancet Oncol. 2006;7:119–26.

    Article  CAS  PubMed  Google Scholar 

  51. Yang XF, Ge YM, Zhang HT, Ning HM, Jiang JQ, Qi YH, Wang ZL. Damaging effects of water-borne cadmium chloride on DNA of lung cells of immature mice. Genet Mol Res. 2012;11:4323–9.

    Article  CAS  PubMed  Google Scholar 

  52. Adams SV, Quraishi SM, Shafer MM, Passarelli MN, Freney EP, Chlebowski RT, Luo JH, Meliker JR, Mu L, Neuhouser ML, Newcomb PA. Dietary cadmium exposure and risk of breast, endometrial, and ovarian cancer in the Women’s Health Initiative. Environ Health Perspect. 2014;122:594–600.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  53. Liang ZL, Wu DD, Yao Y, Yu FY, Yang L, Tan HW, Hylkema MN, Rots MG, Xu YM, Lau ATY. Epiproteome profiling of cadmium-transformed human bronchial epithelial cells by quantitative histone post-translational modification-enzyme-linked immunosorbent assay. J Appl Toxicol. 2018;38:888–95.

    Article  CAS  PubMed  Google Scholar 

  54. Wu DD, Xu YM, Chen DJ, Liang ZL, Chen XL, Hylkema MN, Rots MG, Li SQ, Lau ATY. Ubiquitin carboxyl-terminal hydrolase isozyme L1/UCHL1 suppresses epithelial-mesenchymal transition and is under-expressed in cadmium-transformed human bronchial epithelial cells. Cell Biol Toxicol. 2021;37:497–513.

    Article  CAS  PubMed  Google Scholar 

  55. Tanwar VS, Zhang X, Jagannathan L, Jose CC, Cuddapah S. Cadmium exposure upregulates SNAIL through miR-30 repression in human lung epithelial cells. Toxicol Appl Pharmacol. 2019;373:1–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Vanlaeys A, Fouquet G, Kischel P, Hague F, Pasco-Brassart S, Lefebvre T, Rybarczyk P, Dhennin-Duthille I, Brassart B, Ouadid-Ahidouch H, et al. Cadmium exposure enhances cell migration and invasion through modulated TRPM7 channel expression. Arch Toxicol. 2020;94:735–47.

    Article  CAS  PubMed  Google Scholar 

  57. Wei Z, Shan Z, Shaikh ZA. Epithelial-mesenchymal transition in breast epithelial cells treated with cadmium and the role of Snail. Toxicol Appl Pharmacol. 2018;344:46–55.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. Shan Z, Wei Z, Shaikh ZA. Suppression of ferroportin expression by cadmium stimulates proliferation, EMT, and migration in triple-negative breast cancer cells. Toxicol Appl Pharmacol. 2018;356:36–43.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Shi H, Sun X, Kong A, Ma H, Xie Y, Cheng D, Wong CKC, Zhou Y, Gu J. Cadmium induces epithelial-mesenchymal transition and migration of renal cancer cells by increasing PGE2 through a cAMP/PKA-COX2 dependent mechanism. Ecotoxicol Environ Saf. 2021;207:111480.

    Article  CAS  PubMed  Google Scholar 

  60. Fujiki K, Inamura H, Miyayama T, Matsuoka M. Involvement of Notch1 signaling in malignant progression of A549 cells subjected to prolonged cadmium exposure. J Biol Chem. 2017;292:7942–53.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Chakraborty PK, Scharner B, Jurasovic J, Messner B, Bernhard D, Thévenod F. Chronic cadmium exposure induces transcriptional activation of the Wnt pathway and upregulation of epithelial-to-mesenchymal transition markers in mouse kidney. Toxicol Lett. 2010;198:69–76.

    Article  CAS  PubMed  Google Scholar 

  62. Rocco SA, Koneva L, Middleton LYM, Thong T, Solanki S, Karram S, Nambunmee K, Harris C, Rozek LS, Sartor MA, et al. Cadmium exposure inhibits branching morphogenesis and causes alterations consistent with HIF-1α inhibition in human primary breast organoids. Toxicol Sci. 2018;164:592–602.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Jensen AA, Tuchsen F. Cobalt exposure and cancer risk. Crit Rev Toxicol. 1990;20:427–37.

    Article  CAS  PubMed  Google Scholar 

  64. Rani A, Prasad S. CoCl2-induced biochemical hypoxia down regulates activities and expression of super oxide dismutase and catalase in cerebral cortex of mice. Neurochem Res. 2014;39:1787–96.

    Article  CAS  PubMed  Google Scholar 

  65. Yang G, Xu S, Peng L, Li H, Zhao Y, Hu Y. The hypoxia-mimetic agent CoCl2 induces chemotherapy resistance in LOVO colorectal cancer cells. Mol Med Rep. 2016;13:2583–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  66. Chen DW, Wang H, Bao YF, Xie K. Notch signaling molecule is involved in the invasion of MiaPaCa2 cells induced by CoCl2 via regulating epithelial-mesenchymal transition. Mol Med Rep. 2018;17:4965–72.

    CAS  PubMed  PubMed Central  Google Scholar 

  67. Cui Y, Li YY, Li J, Zhang HY, Wang F, Bai X, Li SS. STAT3 regulates hypoxia-induced epithelial mesenchymal transition in oesophageal squamous cell cancer. Oncol Rep. 2016;36:108–16.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Huang M, Wang L, Chen J, Bai M, Zhou C, Liu S, Lin Q. Regulation of COX-2 expression and epithelial-to-mesenchymal transition by hypoxia-inducible factor-1α is associated with poor prognosis in hepatocellular carcinoma patients post TACE surgery. Int J Oncol. 2016;48:2144–54.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Jin X, Luan H, Chai H, Yan L, Zhang J, Wang Q, Cao L. Netrin-1 interference potentiates epithelial-to-mesenchymal transition through the PI3K/AKT pathway under the hypoxic microenvironment conditions of non-small cell lung cancer. Int J Oncol. 2019;54:1457–65.

    CAS  PubMed  Google Scholar 

  70. Chu CY, Jin YT, Zhang W, Yu J, Yang HP, Wang HY, Zhang ZJ, Liu XP, Zou Q. CA IX is upregulated in CoCl2-induced hypoxia and associated with cell invasive potential and a poor prognosis of breast cancer. Int J Oncol. 2016;48:271–80.

    Article  CAS  PubMed  Google Scholar 

  71. Fei F, Zhang D, Yang Z, Wang S, Wang X, Wu Z, Wu Q, Zhang S. The number of polyploid giant cancer cells and epithelial-mesenchymal transition-related proteins are associated with invasion and metastasis in human breast cancer. J Exp Clin Cancer Res. 2015;34:158.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  72. Lester RD, Jo M, Montel V, Takimoto S, Gonias SL. uPAR induces epithelial-mesenchymal transition in hypoxic breast cancer cells. J Cell Biol. 2007;178:425–36.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  73. Thongchot S, Loilome W, Yongvanit P, Dokduang H, Thanan R, Techasen A, Namwat N. Chloroquine exerts anti-metastatic activities under hypoxic conditions in cholangiocarcinoma cells. Asian Pac J Cancer Prev. 2015;16:2031–5.

    Article  PubMed  Google Scholar 

  74. Kong D, Zhang F, Shao J, Wu L, Zhang X, Chen L, Lu Y, Zheng S. Curcumin inhibits cobalt chloride-induced epithelial-to-mesenchymal transition associated with interference with TGF-β/Smad signaling in hepatocytes. Lab Invest. 2015;95:1234–45.

    Article  CAS  PubMed  Google Scholar 

  75. Liu L, Xiao W. Notch1 signaling induces epithelial-mesenchymal transition in lens epithelium cells during hypoxia. BMC Ophthalmol. 2017;17:135.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  76. Jacobs JA, Testa MS. Overview of chromium(VI) in the environment: background and history. 2005. https://doi.org/10.1201/9780203487969.ch1

  77. Langård S, Vigander T. Occurrence of lung cancer in workers producing chromium pigments. Br J Ind Med. 1983;40:71–4.

    PubMed  PubMed Central  Google Scholar 

  78. Guertin J. Toxicity and Health Effects of Chromium (All Oxidation States): Chromium(VI) Handbook; 2004. https://doi.org/10.1201/9780203487969

  79. Goyer RA, Clarkson TW. Toxic effects of metals. In: Klaasen CD, editor. Casarett and doullis toxicology: the basic science of poisons. 6th ed. New York: McGraw-Hill; 2001. p. 861–7.

    Google Scholar 

  80. Zhang C, Cai K, Feng Q, Xu Y, Zhang Z. Chromium(VI) promotes cell migration through targeting epithelial-mesenchymal transition in prostate cancer. Toxicol Lett. 2019;300:10–7.

    Article  CAS  PubMed  Google Scholar 

  81. Ding SZ, Yang YX, Li XL, Michelli-Rivera A, Han SY, Wang L, Pratheeshkumar P, Wang X, Lu J, Yin YQ, et al. Epithelial-mesenchymal transition during oncogenic transformation induced by hexavalent chromium involves reactive oxygen species-dependent mechanism in lung epithelial cells. Toxicol Appl Pharmacol. 2013;269:61–71.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  82. Li WJ, Yang CL, Chow KC, Kuo TW. Hexavalent chromium induces expression of mesenchymal and stem cell markers in renal epithelial cells. Mol Carcinog. 2016;55:182–92.

    Article  PubMed  CAS  Google Scholar 

  83. Genchi G, Carocci A, Lauria G, Sinicropi MS, Catalano A. Nickel: human health and environmental toxicology. Int J Environ Res Public Health. 2020;17:679.

    Article  CAS  PubMed Central  Google Scholar 

  84. Seilkop SK, Oller AR. Respiratory cancer risks associated with low-level nickel exposure: an integrated assessment based on animal, epidemiological, and mechanistic data. Regul Toxicol Pharmacol. 2003;37:173–90.

    Article  CAS  PubMed  Google Scholar 

  85. Genchi G, Carocci A, Lauria G, Sinicropi MS, Catalano A. Nickel: human healthand environmental toxicology. Int J Environ Res Public Health. 2020;17:679.

    Article  CAS  PubMed Central  Google Scholar 

  86. Wu CH, Hsiao YM, Yeh KT, Tsou TC, Chen CY, Wu MF, Ko JL. Upregulation of microRNA-4417 and its target genes contribute to nickel chloride-promoted lung epithelial cll fibrogenesis and tumorigenesis. Sci Rep. 2017;7:15320.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  87. Jose CC, Jagannathan L, Tanwar VS, Zhang X, Zang C, Cuddapah S. Nickel exposure induces persistent mesenchymal phenotype in human lung epithelial cells through epigenetic activation of ZEB1. Mol Carcinog. 2018;57:794–806.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  88. Hu GF. Copper stimulates proliferation of human endothelial cells under culture. J Cell Biochem. 1998;69:326–35.

    Article  CAS  PubMed  Google Scholar 

  89. Scappaticci FA. Mechanisms and future directions for angiogenesis-based cancer therapies. J Clin Oncol. 2002;20:3906–27.

    Article  CAS  PubMed  Google Scholar 

  90. Li S, Zhang J, Yang H, Wu C, Dang X, Liu Y. Copper depletion inhibits CoCl2-induced aggressive phenotype of MCF-7 cells via downregulation of HIF-1 and inhibition of Snail/Twist-mediated epithelial-mesenchymal transition. Sci Rep. 2015;5:12410.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  91. Li Q, Chen H, Huang X, Costa M. Effects of 12 metal ions on iron regulatory protein 1 (IRP-1) and hypoxia-inducible factor-1 alpha (HIF-1alpha) and HIF-regulated genes. Toxicol Appl Pharmacol. 2006;213:245–55.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  92. Martin F, Linden T, Katschinski DM, Oehme F, Wenger RHJB. Copper-dependent activation of hypoxia-inducible factor (HIF)-1. Blood. 2005;105:4613–9.

    Article  CAS  PubMed  Google Scholar 

  93. Guo H, Jian Z, Liu H, Cui H, Deng H, Fang J, Zuo Z, Wang X, Zhao L, Geng Y, Ouyang P, Tang H. TGF-β1-induced EMT activation via both Smad-dependent and MAPK signaling pathways in Cu-induced pulmonary fibrosis. Toxicol Appl Pharmacol. 2021;418:115500.

    Article  CAS  PubMed  Google Scholar 

  94. Sauna ZE, Shukla S, Ambudkar SV. Disulfiram, an old drug with new potential therapeutic uses for human cancers and fungal infections. Mol Biosyst. 2005;1:127–34.

    Article  CAS  PubMed  Google Scholar 

  95. Li Y, Fu SY, Wang LH, Wang FY, Wang NN, Cao Q, Wang YT, Yang JY, Wu CF. Copper improves the anti-angiogenic activity of disulfiram through the EGFR/Src/VEGF pathway in gliomas. Cancer Lett. 2015;369:86–96.

    Article  CAS  PubMed  Google Scholar 

  96. Li Y, Wang LH, Zhang HT, Wang YT, Liu S, Zhou WL, Yuan XZ, Li TY, Wu CF, Yang JY. Disulfiram combined with copper inhibits metastasis and epithelial-mesenchymal transition in hepatocellular carcinoma through the NF-κB and TGF-β pathways. J Cell Mol Med. 2018;22:439–51.

    Article  CAS  PubMed  Google Scholar 

  97. Ojuederie OB, Babalola OO. Microbial and plant-assisted bioremediation of heavy metal polluted environments: A review. Int J Environ Res Public Health. 2017;14:1504.

    Article  PubMed Central  CAS  Google Scholar 

  98. Dixit R, Wasiullah, Malaviya D, Pandiyan K, Singh UB, Sahu A, Shukla R, Singh BP, Rai JP, Sharma PK et al. Bioremediation of heavy metals from soil and aquatic environment: An overview of principles and criteria of fundamental processes. Sustainability. 2015;7:2189–212.

  99. Tak HI, Ahmad F, Babalola OO. Advances in the application of plant growth-promoting rhizobacteria in phytoremediation of heavy metals. Rev Environ Contam Toxicol. 2013;223:33–52.

    CAS  PubMed  Google Scholar 

  100. Ayangbenro AS, Babalola OO. A New strategy for heavy metal polluted environments: a review of microbial biosorbents. Int J Environ Res Public Health. 2017;14:94.

    Article  PubMed Central  CAS  Google Scholar 

Download references

Acknowledgements

We would like to thank members of the Lau And Xu laboratory for critical reading of this manuscript.

Funding

This work was supported by the grants from the National Natural Science Foundation of China (31771582 and 31271445), the Guangdong Natural Science Foundation of China (2017A030313131), the “Thousand, Hundred, and Ten” Project of the Department of Education of Guangdong Province of China, the Basic and Applied Research Major Projects of Guangdong Province of China (2017KZDXM035 and 2018KZDXM036), the “Yang Fan” Project of Guangdong Province of China (Andy T. Y. Lau-2016 and Yan-Ming Xu-2015), and the Shantou Medical Health Science and Technology Plan (200624165260857).

Author information

Authors and Affiliations

Authors

Contributions

Writing—original draft preparation, X-LC, Y-MX and ATYL; writing—review and editing, X-LC, Y-MX and ATYL; supervision, Y-MX and ATYL; funding acquisition, Y-MX and ATYL. All authors read and approved the final manuscript.

Corresponding author

Correspondence to Andy T. Y. Lau.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

The authors declare that they have no competing interests.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Chen, XL., Xu, YM. & Lau, A.T.Y. Toxic metals in the regulation of epithelial–mesenchymal plasticity: demons or angels?. Cancer Cell Int 22, 237 (2022). https://doi.org/10.1186/s12935-022-02638-3

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12935-022-02638-3

Keywords