Open access peer-reviewed chapter

Causes and Pathogenesis of Malignant Mesothelioma

Written By

Evdoxia Gogou, Sotirios G. Zarogiannis, Dimitra Siachpazidou, Chryssi Hatzoglou and Konstantinos I. Gourgoulianis

Submitted: 21 January 2022 Reviewed: 11 February 2022 Published: 17 March 2022

DOI: 10.5772/intechopen.103669

From the Edited Volume

Mesothelioma - Diagnostics, Treatment and Basic Research

Edited by Ilze Strumfa

Chapter metrics overview

267 Chapter Downloads

View Full Metrics

Abstract

Malignant mesothelioma (MM) is a malignancy that arises from the mesothelium, a thin layer of tissue that covers the body’s serous cavities, such as the pleural, peritoneal, pericardial, and tunica vaginalis of the testis. More than 80% of all mesothelioma cases originate from the pleura and approximately 75–80% of patients are males. It is almost always fatal with most of those affected dying within a year of diagnosis. Asbestos exposure is the most common cause of MM, which mostly affects the pleura. Various factors, including other mineral fibers, carbon nanotubes, or genetic mutations, are also suggested to have a role in the development of MM. The involvement of asbestos, other mineral fibers, nanotechnological products, the simian virus SV40, ionizing radiation, genetic factors, and inflammation in the development of MM has been discussed in this chapter. This study focuses on the role of other mineral fibers, such as erionite, fluoroedenite, balangeroite, and carbon nanotubes, as well as genetic mutations in BAP1 and other genes, in the pathogenesis of MM. The etiology of MM is considered to be complex, and greater knowledge of the pathogenetic pathways may lead to the identification of effective and personalized treatment targets.

Keywords

  • causes of mesothelioma
  • pathogenesis of mesothelioma
  • asbestos
  • BAP1 mutations
  • carbon nanotubes
  • mineral fibers

1. Introduction

Malignant mesothelioma (MM) is a rare and aggressive cancer that affects the mesothelial cells lining the serosal membranes of body cavities, such as the pleura (83% of cases), peritoneum (11%), pericardium, and tunica vaginalis (1–2%) [1, 2, 3]. MM is histologically classified into three types—epithelioid accounting for 80% of the cases, sarcomatoid accounting for more than 10%, and biphasic, which has both epithelioid and sarcomatoid features [4, 5]. The epithelioid subtype is associated with a better prognosis compared to sarcomatoid and biphasic subtypes [5, 6]. Histology and TNM (tumor lymph nodes metastasis) staging are the main prognostic factors and the prognosis remains poor with a median survival from 4 to 19 months [5].

A total of 80% of MM cases concern the pleura and the main cause is asbestos exposure [1]. Approximately 50% of patients with MM have a history of prior asbestos exposure [7]. The median age of diagnosis is 75 years of age and the latency period (the period from the initial asbestos exposure until the diagnosis of mesothelioma) is around 30–40 years [8]. The incidence of mesothelioma is still increasing, despite the wide prohibition of asbestos use. Except for asbestos, exposure to other mineral fibers having similar characteristics, such as erionite or fluoro-edenite, has been implicated in the development of MM [1]. A limited number of MMs are attributed to exposure to ionizing radiation for diagnostic or therapeutic purposes. Asbestos has been widely used for decades globally and 10–17% of those highly exposed to asbestos develop MM [9]. This observation has led to the hypothesis that a possible role of genetic risk factors modifies the effect of asbestos exposure [1]. Recent studies have suggested germline mutations in DNA repairs genes, such as BAP1 (BRCA-1-associated protein) in patients with pleural MM [10, 11]. Approximately 21–63% of MMs involve BAP1 somatic or germline mutations, while 22% of patients with BAP1 mutations will develop MM at some point [2].

During the last years, there have been advances in the understanding of the biology and pathogenesis of mesothelioma. The pathogenesis of MM is thought to be multifactorial and a better understanding of the pathogenetic mechanisms may enable the identification of efficient and personalized treatment patterns for precision medicine. The purpose of our study is to present the causes of mesothelioma by enriching them with the latest data and also describe the possible pathogenetic mechanisms for the development of this insidious cancer.

Advertisement

2. Causes of malignant mesothelioma

2.1 Asbestos exposure

The main cause of MM is exposure to air-born asbestos [11, 12]. Asbestos is a silicate mineral classified into two major groups—the amphiboles group that are sharp, needle-like fibers including crocidolite (known as blue asbestos), amosite (brown asbestos), tremolite, actinolite, and anthophyllite, and the serpentines group that are curly fibers, including chrysotile (known as white asbestos) [12, 13, 14, 15]. All asbestos fibers are considered as carcinogenic by the WHO and the International Agent for Research on Cancer (IARC) (group 1) [12, 16]. The latency period varies between 20 and 70 years [17, 18].

Asbestos-related mesothelioma cases vary by gender, anatomical region, fiber type, and occupation [19, 20, 21]. Most pleural MMs in males are caused by occupational amphibole asbestos exposure. From 2 to 18% of those who were heavily occupationally exposed to amphibole, have developed pleural MM. While the incidence of pleural MM among those occupationally exposed to chrysotile ranges from 0% to 0.47% [19]. Peritoneal MM cases have been reported in those with commercial exposure to amphibole asbestos [22]. However, more recent studies reported that almost 50% of persons with peritoneal MM have fiber load within control values indicating a possible other cause in these tumors [19]. Few studies referred to pericardial or testicular MM and their data did not support the role of asbestos in these sites [23].

The risk of developing MM is related to the type of fiber, the severity, and the duration of exposure [17, 24]. The carcinogenic potency of mineral fibers is determined by the dimensions, durability, dose, and physical properties. Bioavailability after inhalation is affected by fiber dimensions, durability, and dose. Long and thin fibers are associated with higher cytotoxicity and mutagenesis [17, 25]. The WHO distinguished asbestos fibers in short asbestos fibers (SAF) with length < 5 μm and long asbestos fibers (LAF) with length > 5 μm, diameter < 3 μm, and length/diameter ratio > 3 [11, 26]. The longer asbestos fibers, the more carcinogenic potential [26]. Furthermore, fiber biopersistence influences tumorigenesis. The shorter biopersistence, the lower carcinogenic potential as observed in serpentine chrysotile compared with amphiboles and erionite [17]. However, if the exposure to short biopersistence fibers, such as chrysotile, is prolonged, the mesothelial cells could be transformed [27]. MM from occupational asbestos exposure is mainly caused by crocidolite and amosite fibers [28, 29].

The mass mining of asbestos began in the twentieth century and it was mainly used for insulation against heat, fire, and corrosion, while its previous use was in pottery [12]. Thus, high-risk occupations include engineers who work on brake and clutch lining, builders, dockyard and shipyard workers, plumbers, and electricians [4].

Asbestos exposure can occur mainly occupationally for asbestos workers and nonoccupationally including domestic, neighborhood, or environmental exposure [4, 12, 30]. The risk of developing pleural MM after nonoccupational exposure depends on the types of fibers. The risk is greater to amphiboles exposure than to chrysotile [11]. However, it is complicated to establish accurately asbestos exposure. There is unquestionable asbestos exposure in asbestos miners and shipyards workers. Other categories of workers may not correctly remember events that occurred 30–50 years earlier, as the latency period is too long. Specific questionnaires were developed to identify different levels of exposure within occupational asbestos exposure individuals. Another evidence of asbestos exposure is the measurement of lung tissue fiber content, but this is rarely performed as it is invasive, costly, and for legal implications [11, 12]. The combination of a complete occupational history and radiological evidence of exposure, such as bilateral calcified pleural plaques and/or histological evidence of asbestos fibers in lung tissue, could be used to estimate asbestos exposure. For example, pleural plaques were found in 88% of asbestos exposure patients with mesothelioma [11, 29].

Asbestos use was prohibited in most western countries between 1970 and 2005, except for the USA, where it was only partly banned, and Canada, where the asbestos ban was implemented in 2018 [10, 31]. These countries represent 16% of the world’s population [12]. Unfortunately, in developing countries, asbestos use and mining are ongoing with an annual worldwide production of about 2.2 million metric tons [32]. Hence, the incidence of MM will continue to increase worldwide [12].

2.1.1 Carcinogenic mechanisms of asbestos

When asbestos and other fibers enter the pleura and peritoneum via lymphatics, they reside there for months or years, triggering a chronic inflammatory response stimulated by high mobility group protein B1 (HMGB1) secretion and related inflammasome activation in mesothelial cells, which activates the nuclear factor kappa-light-chain-enhancer of activated B cells (NF-kB) and phosphatidylinositol 3-kinase (PI3K) [33, 34, 35, 36]. This environment promotes the proliferation of mesothelial cells that have spontaneously acquired mutations or are exposed to mutagenic reactive oxygen species generated by inflammatory cells in the area of asbestos deposits [34, 35]. Asbestos-activated macrophages produce reactive oxygen species (ROS), which can cause DNA damage by forming 8-hydroxy-2-deoxyguanosine (8-OHDG) adducts [34]. Ferroptosis, a non-apoptotic, iron-dependent cell death, has recently been linked to asbestos-related carcinogenesis [37]. Hepatocyte growth factor (HGF) may also play a role in asbestos-induced carcinogenesis by activating the PI3K/MEK5/Fra-1 axis (phosphatidylinositol 3-kinase/mitogen/extracellular signal-regulated kinase kinases 5/(Fos-related antigen 1) [38]. Crocidolite and erionite have a longer biopersistence than chrysotile, which explains their greater pathogenicity [35].

Asbestos fibers are phagocytosed by human mesothelial cells, and once inside the cell, they can mechanically interfere with the cell spindle during mitosis, causing chromosomal mutations responsible for carcinogenesis, but this hypothesis has been ruled out [11, 39].

The carcinogenesis mechanism in pleural MM is complex. Inhaled asbestos fibers move to the pleura. Fibers in the pleural space irritate the tissue, resulting in a cycle of tissue injury and repair. When asbestos fibers are phagocytosed by macrophages, oxygen-free radicals are produced, causing intracellular DNA damage and aberrant repair [40]. Asbestos fibers also enter mesothelial cells, interfering with mitosis, causing DNA mutations, and changing chromosome structure. Inflammatory cytokines are released by asbestos-exposed mesothelial cells, including tumor growth factor, platelet-derived growth factor, and vascular endothelial growth factor (VEGF) [40]. This creates an ideal environment for tumor development. Finally, asbestos increases the expression of proto-oncogenes and promotes aberrant cellular proliferation by phosphorylating different protein kinases (mitogen-activated protein and extracellular signal-regulated kinases 1 and 2) [27]. Asbestos fibers are known to cause DNA damage, which is repaired by homologous recombination (HR) and double-strand break repair, mismatch repair system (MMR), and nonhomologous end-joining (NHE) or nucleotide excision repair (NER), putting people with DNA repair faults at a higher risk of developing MM [1, 41, 42, 43, 44].

2.2 Erionite

Erionite is a fibrous type of zeolite and according to its physical characteristics, it resembles amphiboles amosite or crocidolite [19, 45]. Chemically, it consists of potassium aluminum silicate with various amounts of calcium and sodium [19]. Deposits of erionite have been described in the Cappadocian region of Turkey, in the intermountain west of the United States from Oregon into Mexico and the Sierra Madre region [46, 47]. High amounts of airborne erionite were found in North Dakota, where miles of roads were surfaced with erionite-containing gravel [48]. Also, erionite has been identified in North-Eastern Italy [49].

Studies have shown that erionite is a carcinogenic fiber that causes the MM epidemic in some Cappadocian villages in Turkey [48, 50]. There, erionite was used to build houses and pave roads. Environmental exposure to erionite fibers was documented not only in Cappadocian of Turkey but also in Mexico, North Dakota, Nevada, and California [46, 47, 48, 51]. More specifically, in North Dakota erionite has been used to pave roads, in Nevada referred exposure to asbestos, erionite, and other types of fibers, and in California referred exposure to chrysotile and tremolite [48, 51, 52]. These fibers get released into the air due to human activities, such as mining, road construction, and off-road driving. Environmental exposure often begins at birth and occurs randomly among males and females. That is why MM caused by environmental exposure tends to occur at a younger age < 55 years with a 1:1 male: female, contrary to MM caused by occupational asbestos exposure that has a ratio 3:1 [53]. Mineralogical and pedigree analyses in villages in Cappadocian of Turkey have revealed that in addition to environmental exposure there may also be an autosomal dominant genetic susceptibility to MM [50, 54]. Middle-aged patients diagnosed with mesothelioma in North America reported living in Mexico at a young age and having emigrated. Their fiber burden analysis in lung tissue demonstrated a high aspect ratio of erionite fibers, which existed in high concentrations in Mexico [46, 55]. Experimental animal studies have confirmed the high carcinogenic potential of erionite [56, 57]. Erionite also could cause other disorders, such as mesothelial hyperplasia, dysplasia, and pleural fibrosis [56]. The way erionite is thought to cause MM is by activation of the NLRP3 (NOD-, LRR- and pyrin domain-containing protein 3) inflammasome, which in turn triggers an autocrine feedback loop in mesothelial cells, modulated by the interleukin-1 receptor [58].

2.3 Fluoro-edenite

Fluoro-edenite has similar morphology and composition to actinolite and tremolite. This mineral was extracted from an area located at Southeast of Biancavilla in Catania, Eastern Sicily, Italy. It was used as a building material, for road paving, for residential and commercial plaster, and also for mortar construction [19]. A study showed a 10-fold increase in pleural neoplasms among those exposed to fluoro-edenite [59]. An experimental animal study revealed that when fluoro-edenite was implanted in peritoneal cavities of rats, MM was induced [60]. Fluoro-edenite was classified as Group 1 carcinogenic to humans [61]. According to studies in vitro, fluoro-edenite induces DNA damage and leads to the production of reactive oxygen species (ROS) resulting in decreased cell viability [62].

2.4 Balangeroite

Balangeroite is a fibrous iron-rich magnesium silicate and it is often intergrown with chrysotile deposits. Deposits have been found in Balangero in Italy, after which it is named. It has a similar morphology but lower bio-durability than commercial amphiboles [63, 64]. The role of these amphibole mineral fibers in the induction of MM in Balangero in Italy is controversial with some authors attributing MMs to balangeroite and others blurring its precise role [64, 65]. The controversy arises from the fact that some Balangero chrysotile miners have commercial amphiboles (crocidolite and amosite) in the mineral analysis of lung tissue. Also, it was known that the Balangero mines occasionally milled imported commercial amphibole from South Africa [63, 65].

2.5 Other minerals

Other minerals include man-made vitreous fibers, such as rock wool, slag wool, glass fiber and glass filament. All mineral wools are formed by spinning or drawing molten mineral or rock materials, such as slag and ceramics. These are applied to thermal insulation, filtration, soundproofing, and hydroponic medium [66]. The more biopersistent man-made vitreous fibers were classified by IARC (International Agency for Research on Cancer) as “possible carcinogenic to humans” (group 2b) [66]. More biopersistent refractory ceramic fibers have been linked to the induction of MM in Syrian golden hamsters exposed to high-dose chronic inhalation experiments [67]. Some case reports of MM have been related to beryllium, nickel, and crystalline silica, but these data have not been supported by epidemiological studies [19]. There is an increased risk of pleural MM for those exposed to both asbestos and mineral wool or silica according to one study [68].

2.6 Simian Virus 40 (SV40)

SV40 is a DNA polyomavirus that has been reported as a possible etiologic agent for human MM [69]. Human exposure to SV40 occurred between 1955 and 1963 when inactivated and live anti-polio vaccines were administrated to people in the United States, Canada, Europe, Asia, and Africa [70].

SV40 sequences have been found by the polymerase chain reaction (PCR) method in various human cancers, such as MM, non-Hodgkin lymphoma, osteosarcoma, and thyroid tumors [71]. There are data available on the role of SV40 in the pathogenesis of human MM. For example, this virus activates genes promoting cell progression and proliferation, it induces apoptosis of mesothelial cells transfected with antisense DNA to the SV40, and MM harboring SV40 has a poorer prognosis compared to SV40-negative MM [69]. Mesothelial cells are susceptible to infection and transformation by SV40 [72]. The viral genome encodes oncogenic proteins like large T-antigen (Tag), which inactivate the tumor suppressor activity of p53 and p-retinoblastoma family proteins. However, the presence of SV40 DNA and protein in MM has not led to the definitive causal relationship between the virus and MM development [73]. According to some researchers, SV40 in humans may be a passenger virus in the mesothelial cells without causing pathology or tumorigenesis [69]. Overall, the role of SV40 as an etiologic agent in human MM is still in debate.

2.7 Radiation

Ionizing radiation is a high-risk factor for malignancy development. Its effect is cumulative, so once received, the effects remain in the body for life. Individuals with increased levels of exposure to ionizing radiation have a greater risk of malignancies later in their life [74]. The carcinogenic risk associated with exposure to ionizing radiation has been evaluated previously in the IARC monographs [66].

The evidence linking radiation to MM in humans comes—(i) from clinical studies involving patients who had previously received radiotherapy for tumors, (ii) from reported cases of MM occurring after the use of the contrast agent thorotrast, and (iii) from studies of workers exposed to prolonged lower levels of radiation [19]. Cases of pleural, peritoneal, and pericardial mesothelioma have been reported after radiotherapy in childhood or adulthood due to lymphoma, genital, renal, and breast neoplasms [75, 76]. The radiation-induced MMs had a latent period from 5 to 50 years and an equal male: female ratio [77]. The intravenous thorotrast administration has caused not only MM but also hepatocellular carcinoma, hemangioendothelioma, and cholangiocarcinoma. The radioactive Thorotrast (232ThO2) is insoluble and after injection, deposits in organs and is associated with slow decay and prolonged alpha-ray emission [19]. Cases of MM have been reported in radiation technologists and among workers in the atomic energy industry [78]. A genetic analysis study has shown that radiation-induced MMs have copy number gains outnumbering deletions, which are more common in asbestos-induced MMs, signifying potential different molecular mechanisms of induction [79]. Overall, radiation is a risk factor to MM in directly irradiated tissues and to a lesser extent in tissues distant from the target site.

2.8 Chronic inflammation and MM

Chronic serosal membranes inflammations can induce MM of pleura and peritoneum [80]. Therapeutic plombage used as a treatment for pulmonary tuberculosis and longstanding chronic empyema could induce pleural MM. Moreover, recurrent peritonitis as a result of relapsing diverticulitis or Crohn’s disease or Familial Mediterranean Fever, ventriculoperitoneal shunts for hydrocephaly have been reported as a cause of peritoneal MM [80, 81]. Chronic interleukin-6 production has been linked to MM pathogenesis as a regulatory cytokine in the acute phase response [19].

2.9 BAP1 (BRCA-1-associated protein 1)

Recently, many researchers are concerned with the role of BAP1 in mesothelioma. BAP1 is a nuclear protein, which is encoded by a tumor suppressor gene (BAP1 gene) located on chromosome 3p21.1 [10, 82]. BAP1 was discovered in 1998 as a novel ubiquitin carboxyl-terminal hydrolase, an enzyme responsible for removing ubiquitin from protein substrates [83]. BAP1 is binding to BRCA1 enhancing its tumor-suppressive activity [7, 83]. Also, BAP1 regulates proteins involved in DNA damage repair, cellular differentiation, chromatin modulation, cell cycle control and cell proliferation, immune regulation, and consequently, it has a tumor-suppressive effect [7]. BAP1 is a nuclear protein that belongs to a family of multiprotein transcriptional regulators that control genes related to metabolism, mitochondrial function, and cell proliferation [33, 37]. The identification of BAP1 as a key regulator of cell death and metabolism aided in the description of the complex set of molecular events mediated by asbestos carcinogenesis [84].

Clinical reports have shown that BAP1 is commonly lost or inactivated in various cancers [85]. An increase in the spontaneous development of breast cancer, lung, ovarian and a few cases of MM that are not related to asbestos in about half of mice with genetically engineered BAP1 mutations that match those found in BAP1 cancer syndrome families supports the idea that BAP1 is a tumor suppressor [86]. BAP1 mutations occur in a wide range of people. According to a study, BAP1 germline mutations were found in 7.7% of spontaneous MM cases [87]. It is suggested that germline mutations in BAP1 are thought to start with just one abnormal allele. Low-level asbestos exposure resulted in second allele mutations in genetically vulnerable hosts, resulting in the development of MM linked to BAP1 [88]. Experiments with animal subjects have backed up the aforementioned theory. Experiments in BAP1+/− mice revealed that after intraperitoneal injection of crocidolite asbestos, animals developed MM at twice the rate of wild-type mice, while no MMs were observed in BAP+/− mice not exposed to asbestos [89]. Other researchers discovered that BAP1 knockout mice developed MM without ever having been exposed to asbestos [86]. Another study presented that none of the patients with MM and gene mutations reported occupational asbestos exposure, highlighting that these tumors were either due to low levels of environmental exposure or not due to exposure to carcinogenic fibers [90]. The same study showed that most patients were female and almost half of the tumors were located in the peritoneum, arguing that they were not related to asbestos exposure, as if the cause was both asbestos exposure and genetic predisposition, then the male: female ratio would be maintained and most tumors would be located in the pleura [90].

Taking into consideration the functional role of BAP1 in many cellular pathways implicated in cancer, it is not surprising that the BAP1 gene is mutated in a variety of tumors [85]. BAP1 mutations observed in cancer are primarily inactivating mutations, such as chromosomal deletions of the BAP1 gene, leading to loss of function. BAP1 mutations occur in both germline and somatic forms [7, 85]. Germline mutations of the BAP1 gene are inherited in an autosomal dominant pattern and constitute a novel tumor predisposition syndrome (BAP1-TPDS) conferring a high risk of hereditary cancers [87, 91]. The cancers associated with this syndrome are MM, uveal or cutaneous melanoma, and renal cell carcinoma [92]. MM is the second most common cancer identified with BAP1-TPDS accounting for 22% of tumors with a median age of onset of 46 years and a seven-fold longer survival rate compared to a patient with sporadic MM [90]. Somatic BAP1 mutations appear in similar types of tumors as in patients with germline mutations. A total of 50% of MM patients were found to have somatic BAP1 mutations and interestingly, they show significantly longer survival than those without mutations on BAP1 [93, 94]. Around 21–63% of MM patients have BAP1 mutations (germline or somatic), and 22% of those who have BAP1 mutations will develop MM. BAP1 genetic mutations are normally present in all cells with one mutant allele, whereas somatic inactivation of a second allele causes cancer [95, 96]. For BAP1 mutation carriers, the gene–environmental interaction is thought to play a key role in cancer susceptibility [95]. In the general population, BAP1 mutations are uncommon, and there are no homozygotes [97]. In distinct cases, their prevalence has been reported to be 1–2% for uveal melanoma, 0.5% for cutaneous melanoma, and 0–7% for MM, rising to 25%, 0.7%, and 20%, respectively, in family cases [43, 44]. In MMs and other malignancies linked to BAP1-TPDS, tumor aggressiveness varies greatly, and the underlying genetic processes are unknown [10].

2.10 Other genes linked to MM pathogenesis

A handful of the genes implicated in chromatin regulation are mutated in MM patients. Some genes [CDKN2A (cyclin-dependent kinase inhibitor 2A gene), TMEM127 (transmembrane protein 127 gene)] encode tumor suppressor proteins involved in cell growth, proliferation, and survival. Other genes, such as NF2 (neurofibromin 2), encode proteins that modulate signaling pathways for modulating cell shape, cell growth, and cell adhesion [10, 98]. Other genes, such as KDR (kinase insert domain receptor), encode vascular endothelial growth factor (VEGF) receptors that increase endothelial cell proliferation, survival, migration, and differentiation [99]. The pathophysiological mechanisms underlying the development of MM as a result of these genetic mutations are still unknown.

Many studies highlight different clinical parameters that can predict the presence of an inherited mutation in MM, such as minimal asbestos exposure, peritoneal disease, young age, and a second cancer diagnosis [42, 43, 44]. This finding is significant because it could lead to the development of clinical panel-based genetic testing and the adoption of clinical genetic testing recommendations. For MM patients, genetic testing would be extremely beneficial because it would allow for the early detection and prevention of malignancies in high-risk individuals. The earlier cancers are detected and treated, the better the chances of survival [10].

According to comprehensive Genome-Wide Association Studies (GWAS) on MM, the most significant SNP (single nuclear polymorphisms) were found in genes involved in cell adhesion, migration, and apoptosis, and may promote carcinogenesis via mechanisms triggered by the human immune system’s response to asbestos fibers [100]. Reactive oxygen species (ROS) and free radicals, which arise as a result of inhaled asbestos fiber, are thought to have a role in asbestos toxicity and carcinogenicity. Genetic polymorphisms in detoxification genes encode proteins that are involved in the detoxification and clearance of ROS or change enzyme function, which may increase cancer risk. Furthermore, genetic polymorphisms in DNA repair genes result in a deficiency in DNA repair pathways, which fail to defend against the oxidative stress generated by asbestos fibers, ultimately leading to an increased risk of carcinogenesis [101]. Individuals who were homozygotes or heterozygotes in one of four DNA repair genes were more likely to develop pleural MM than controls [10, 99, 101].

Reduced expression of critical molecules in the p53 tumor-suppressor gene pathway, such as p14, p16, and NF2-MERLIN (Moesin-ezrin-radixin-like protein), has been discovered as a result of genetic profiling of pleural MM [27, 102]. Other genes, such as BRCA-associated protein 1 (BAP1), set domain containing 2 (SETD2), unc-like autophagy activating kinase (ULAK2), ryanodine receptor 2 (RR2), cilia and flagella associated protein 45 (CFAP45), and set domain bifurcated 1 (SETDB1), have been shown to have deletions or loss mutations in pleural MM [17].

More research is needed to provide a full picture of the genes that predispose to mesothelioma and their role in the molecular pathways of asbestos carcinogenesis that have been revealed, such as chronic inflammation and altered metabolism.

2.11 The role of genes and environment

Carcinogenesis is frequently linked to somatic gene changes that disrupt DNA repair systems, resulting in an accumulation of DNA damage and an increase in the proportion of cells with damaged DNA. Cancer could arise if these cells kept their survival mechanisms. Inherited mutations affecting DNA repair and other genes may exacerbate carcinogenesis by increasing vulnerability to environmental carcinogens [11, 33]. In the subject of carcinogens, the current method is to explore genes and environment interactions by combining genetic and environmental investigations.

2.12 Carbon nanotubes and mesothelioma

Carbon nanotubes are one-dimensional fibrous nanomaterials that resemble asbestos fibers in their physical properties. In 2014, WHO and IARC (International Agency for Research on Cancer) classified long, rigid multiwall carbon nanotubes as group 2B possibly carcinogenic for humans [61, 103]. Only one type of carbon nanotube was categorized in group 2B-possibly carcinogenic to humans in 2014 and this was a commercial product called “MWCNT-7.” (multiwall carbon nanotubes-7) [61, 104, 105]. MWCNT-7 are multiwall carbon nanotubes with a structure comparable to asbestos and are biopersistent [106]. Toxicological investigations in animals showed that some forms of carbon nanotubes can cause MM. In animal experimental models, long, large-diameter, rigid multiwall carbon nanotubes supplied by intraperitoneal or intrascrotal injection or trans-tracheal intrapulmonary spraying were shown to develop MM [106, 107, 108]. Longer, rigid multiwall carbon nanotubes translocated to the parietal pleura, causing more inflammation, fibrosis, and localized mesothelial cell proliferation than shorter, thinner agglomerates [89]. Many animal experimental studies have shown similar findings [106, 109]. Asbestos fibers and multiwall carbon nanotubes have physicochemical similarities and differences [103]. The “fiber pathogenicity paradigm” identifies width, length, biopersistence, and mechanical bending stiffness as predictors of pathogenicity and carcinogenicity in carbon nanotubes, or other fibrous nanomaterials, and metallic nanowires [110].

Near the end of the twentieth century, carbon nanotubes were found and manufactured. Composite materials, thin coatings and films, microelectronics, energy storage, environmental remediation, and nanomedicine are all areas where they are used [111]. The Royal Society and Royal Academy of Engineering acknowledged the physical similarities between carbon nanotubes and asbestos fibers in 2004, as well as the potential for human health hazards [112]. Chemical vapor deposition (CVD) is the most common industrial method for producing carbon nanotubes, with transition metals catalyzing the breakdown of a carbon-containing organic vapor. Carbon nanotubes can be discharged as dry powders during the manufacturing and processing phases [110]. The National Institute of Occupational Safety and Health (NIOSH) in the United States has set a recommended exposure limit of 1 μg/m3 [113]. To prevent repeating the history of asbestos-related disorders, the ultimate goal is to commercialize nanomaterials while simultaneously considering potential human health risks [103].

2.12.1 Pathogenicity of carbon nanotubes

Some forms of carbon nanotubes resemble asbestos fibers in length and rigidity. Long, thin, biopersistent fibers are thought to enter the pleural space, obstruct clearance through lymphatic stomata on the parietal pleura, and cause frustrated phagocytosis, oxidant generation, and persistent inflammation, ultimately leading to MM [110]. Furthermore, the hydrophobicity of raw carbon nanotubes hampered their dispersion in biological conditions, causing them to clump together in rope-like formations or tangled clumps. These clusters clump together to create discrete multifocal granulomas, which comprise macrophage and fibroblast aggregates [114, 115]. Carbon nanotubes were also found in lymph nodes in the mediastinum. Individual carbon nanotubes may be gradually released from pulmonary agglomerates over time and translocated to the lung interstitium, pleura, and regional lymph nodes. Also, many researchers discovered that inhaling well-aerosolized single or multi-walled carbon nanotubes caused persistent inflammation and fibrosis [106]. Very thin carbon nanotubes are more prone to form tangled agglomerates than thicker, rigid multiwall nanotubes. DNA damage can be caused by short multiwall carbon nanotubes with a length of 1 μm [103, 116].

Advertisement

3. Conclusions

Asbestos exposure is the most common cause of MM, however, genetic factors, such as BAP1 gene mutations and exposure to other minerals fibers or nanotechnology products, have also been linked in recent years. It is possible that genetic and environmental factors interact to cause MM development. Knowing the causes of MM can help with early detection and prevention. Furthermore, studying and comprehending the pathogenetic pathways that contribute to the development of mesothelioma can help to find more targeted and effective treatments, hence prolonging survival.

Advertisement

Acknowledgments

The authors would like to thank Chamou Dimitra for the literature assistance.

Advertisement

Competing interests statement

None.

References

  1. 1. Betti M, Aspesi A, Sculco M, Matullo G, Magnani C, Dianzani I. Genetic predisposition for malignant mesothelioma: A concise review. Mutation Research Mutation Research. 2019;781:1-10. Available from: https://www.sciencedirect.com/science/article/pii/S1383574218300656
  2. 2. Kato S, Tomson BN, Buys TPH, Elkin SK, Carter JL, Kurzrock R. Genomic landscape of malignant mesotheliomas. Molecular Cancer Therapeutics. 2016;15(10):2498-2507
  3. 3. Rusch VW, Giroux D, Kennedy C, Ruffini E, Cangir AK, Rice D, et al. Initial analysis of the international association for the study of lung cancer mesothelioma database. Journal of Thoracic Oncology: Official Publication of the International Association for the Study of Lung Cancer. 2012;7(11):1631-1639. DOI: 10.1097/JTO.0b013e31826915f1
  4. 4. Asciak R, George V. Update on biology and management of mesothelioma. European Respiratory Review. 2021;1:1-13. DOI: 10.1183/16000617.0226-2020
  5. 5. Brcic L, Kern I. Clinical significance of histologic subtyping of malignant pleural mesothelioma. Translational Lung Cancer Research. 2020;9(3):924-933
  6. 6. Galateau-salle F, Churg A, Roggli V, Travis WD, Health W. The 2015 World Health Organization classification of tumors of the pleura: Advances since the 2004 classification. Pleural Malignant Mesothelioma. 2015;11(2):142-154
  7. 7. Louie BH, Kurzrock R. BAP1: Not just a BRCA1-associated protein. Cancer Treatment Reviews. 2020;90(August):102091. DOI: 10.1016/j.ctrv.2020.102091
  8. 8. Woolhouse I, Bishop L, Darlison L, De Fonseka D, Edey A, Edwards J, et al. British Thoracic Society Guideline for the investigation and management of malignant pleural mesothelioma. Thorax. 2018;73:i1-i30
  9. 9. Hodgson JT, Darnton A. The quantitative risks of mesothelioma and lung cancer in relation to asbestos exposure. Annals of Occupational Hygiene. 2000;44(8):565-601
  10. 10. Panou V, Røe OD. Inherited genetic mutations and polymorphisms in malignant mesothelioma: A comprehensive review. International Journal of Molecular Sciences. 2020;21(12):1-18
  11. 11. Carbone M, Prasad S, PSH A, Richard Alexander HR, Baas P, Fabrizio Bardelli F, et al. Mesothelioma: Scientific clues for prevention, diagnosis, and therapy. CA: A Cancer Journal for Clinicians. 2019;69(5):402-429
  12. 12. Røe OD, Stella GM. Malignant pleural mesothelioma: History, controversy and future of a manmade epidemic. European Respiratory Society. 2015;24(135):115-131. DOI: 10.1183/09059180.00007014
  13. 13. Stayner L, Welch LS, Lemen R. The worldwide pandemic of asbestos-related diseases. Annual Review of Public Health. 2013;34:205-216
  14. 14. Kumagai-Takei N, Lee S, Srinivas B, Shimizu Y, Sada N, Yoshitome K, et al. The effects of asbestos fibers on human T cells. International Journal of Molecular Sciences. 2020;21(19):1-13
  15. 15. Sporn TA. Mineralogy of asbestos. In: Tannapfel A, editor. Malignant Mesothelioma Recent Results in Cancer Research. Berlin, Heidelberg: Springer; 189, 2011. DOI: 10.1007/978-3-642-10862-4_1
  16. 16. Arsenic IARC, Metals F. Dusts. IARC Monographs on the Evaluation of Carcinogenic Risks to Humans. 2012;100:147-167
  17. 17. Bibby AC, Tsim S, Kanellakis N, Ball H, Talbot DC, Blyth KG, et al. Malignant pleural mesothelioma: An update on investigation, diagnosis and treatment. European Respiratory Society. 2016;25(142):472-486. DOI: 10.1183/16000617.0063-2016
  18. 18. Robinson BM. Malignant pleural mesothelioma: An epidemiological perspective. The Annals of Cardiothoracic Surgery. 2012;1(4):491-496
  19. 19. Attanoos RL, Churg A, Galateau-Salle F, Gibbs AR, Roggli VL. Malignant mesothelioma and its non-asbestos causes. Archives of Pathology & Laboratory Medicine. 2018;142(6):753-760
  20. 20. Spirtas R, Heineman EF, Bernstein L, Beebe GW, Keehn RJ, Stark A, et al. Malignant mesothelioma: Attributable risk of asbestos exposure. Occupational and Environmental Medicine. 1994;51(12):804-811
  21. 21. Rake C, Gilham C, Hatch J, Darnton A, Hodgson J, Peto J. Occupational, domestic and environmental mesothelioma risks in the British population: A case-control study. British Journal of Cancer. 2009;100(7):1175-1183
  22. 22. Browne K, Smither WJ. Asbestos-related mesothelioma: Factors discriminating between pleural and peritoneal sites. British Journal of Industrial Medicine. 1983;40(2):145-152
  23. 23. Lowry SJ, Weiss NS. Geographic distribution of incidence of pericardial and paratesticular mesotheliomas in the USA. Cancer Causes Control. 2016;27:1487-1489. DOI: 10.1007/s10552-016-0825-3
  24. 24. Wiggins J, Brilton MG, Darlison L, Entwisle J, Greenstone MA, Hetzel M, et al. BTS statement on malignant mesothelioma in the UK, 2007. Thorax. 2007;62(Suppl. 2):1-19
  25. 25. Beckett P, Edwards J, Fennell D, Hubbard R, Woolhouse I, Peake MD. Demographics, management and survival of patients with malignant pleural mesothelioma in the National Lung Cancer Audit in England and Wales. Lung Cancer. 2015;88(3):344-348. DOI: 10.1016/j.lungcan.2015.03.005
  26. 26. Pierce JS, McKinley MA, Paustenbach DJ, Finley BL. An evaluation of reported no-effect chrysotile asbestos exposures for lung cancer and mesothelioma. Critical Reviews in Toxicology. 2008;38(3):191-214
  27. 27. Robinson BWS, Lake RA. Advances in malignant mesothelioma. The New England Journal of Medicine. 2005;353(15):1591-1603
  28. 28. Sluis-Cremer GK, Liddell FDK, Logan WPD, Bezuidenhout BN. The mortality of amphibole miners in South Africa, 1946-80. British Journal of Industrial Medicine. 1992;49(8):566-575
  29. 29. Dodson RF, O’Sullivan M, Corn CJ, McLarty JW, Hammar SP. Analysis of asbestos fiber burden in lung tissue from mesothelioma patients. Ultrastructural Pathology. 1997;21(4):321-336
  30. 30. Panou V, Vyberg M, Meristoudis C, Hansen J, Bøgsted M, Omland Ø, et al. Non-occupational exposure to asbestos is the main cause of malignant mesothelioma in women in North Jutland, Denmark. Scandinavian Journal of Work, Environment & Health. 2019;45(1):82-89
  31. 31. Ruff K. How Canada changed from exporting asbestos to banning asbestos: The challenges that had to be overcome. International Journal of Environmental Research and Public Health. 2017;14(10):1135
  32. 32. Hashim D, Boffetta P. Occupational and environmental exposures and cancers in developing countries. Annals of Global Health. 2014;80(5):393-411. DOI: 10.1016/j.aogh.2014.10.002
  33. 33. Gaudino G, Xue J, Yang H. How asbestos and other fibers cause mesothelioma. Translational Lung Cancer Research. 2020;9(I):S39-S46
  34. 34. Xu A, Wu LJ, Santella RM, Hei TK. Role of oxyradicals in mutagenicity and DNA damage induced by crocidolite asbestos in mammalian cells. Cancer Research. 1999;59(23):5922-5926
  35. 35. Qi F, Okimoto G, Jube S, Napolitano A, Pass HI, Laczko R, et al. Continuous exposure to chrysotile asbestos can cause transformation of human mesothelial cells via HMGB1 and TNF-α signaling. The American Journal of Pathology. 2013;183(5):1654-1666. DOI: 10.1016/j.ajpath.2013.07.029
  36. 36. Carbone M, Yang H. Targeting mechanisms of asbestos and erionite carcinogenesis in mesothelioma. Clinical Cancer Research. 2012;18(3):598-604
  37. 37. Toyokuni S. Iron addiction with ferroptosis-resistance in asbestos-induced mesothelial carcinogenesis: Toward the era of mesothelioma prevention. Free Radical Biology and Medicine. 2019;133:206-215. DOI: 10.1016/j.freeradbiomed.2018.10.401
  38. 38. Ramos-Nino ME, Blumen SR, Sabo-Attwood T, Pass H, Carbone M, Testa JR, et al. HGF mediates cell proliferation of human mesothelioma cells through a PI3K/MEK5/Fra-1 pathway. American Journal of Respiratory Cell and Molecular Biology. 2008;38(2):209-217
  39. 39. Olofsson KMJ. Specificity of asbestos-induced chromosomal aberrations in short-term cultured human mesothelial cells. Cancer Genetics and Cytogenetics. 1989;41(1):33-39. DOI: 10.1016/0165-4608(89)90105-2
  40. 40. Sekido Y. Molecular pathogenesis of malignant mesothelioma. Carcinogenesis. 2013;34(7):1413-1419
  41. 41. Toumpanakis DTSE. DNA repair systems in malignant mesothelioma. Cancer Letter. 2011;312(2):143-149. DOI: 10.1016/j.canlet.2011.08.021
  42. 42. Pastorino S, Yoshikawa Y, Pass HI, Emi M, Nasu M, Pagano I, et al. A subset of mesotheliomas with improved survival occurring in carriers of BAP1 and other germline mutations. Journal of Clinical Oncology. 2018;36(35):3485-3494
  43. 43. Panou V, Gadiraju M, Wolin A, Weipert CM, Skarda E, Husain AN, et al. Frequency of germline mutations in cancer susceptibility genes in malignant mesothelioma. Journal of Clinical Oncology. 2018;36(28):2863-2871
  44. 44. Hassan R, Morrow B, Thomas A, Walsh T, Lee MK, Gulsuner S, et al. Inherited predisposition to malignant mesothelioma and overall survival following platinum chemotherapy. Proceedings of the National Academy of Sciences of the United States of America. 2019;116(18):9008-9013
  45. 45. Dogan AU, Dogan M, Hoskins JA. Erionite series minerals: Mineralogical and carcinogenic properties. Environmental Geochemistry and Health. 2008;30(4):367-381
  46. 46. Ortega-Guerrero MA, Carrasco-Núñez G. Environmental occurrence, origin, physical and geochemical properties, and carcinogenic potential of erionite near San Miguel de Allende, Mexico. Environmental Geochemistry and Health. 2014;36(3):517-529
  47. 47. Baris YI, Grandjean P. Prospective study of mesothelioma mortality in Turkish villages with exposure to fibrous zeolite. Journal of the National Cancer Institute. 2006;98(6):414-417
  48. 48. Carbone M, Baris YI, Bertino P, Brass B, Comertpay S, Dogan AU, et al. Erionite exposure in North Dakota and Turkish villages with mesothelioma. Proceedings of the National Academy of Sciences of the United States of America. 2011;108(33):13618-13623
  49. 49. Giordani Μ, Mattioli Μ, Ballirano P, Pacella A, Cenni M, Boscardin MVL. Geological occurrence, mineralogical characterization, and risk assessment of potentially carcinogenic erionite in Italy. Journal of Toxicology and Environmental Health, Part B. 2017;20(2):81-103
  50. 50. Metintas M, Hillerdal G, Metintas S. Malignant mesothelioma due to environmental exposure to erionite: Follow-up of a Turkish emigrant cohort. The European Respiratory Journal. 1999;13(3):523-526
  51. 51. Pan XL, Day HW, Wang W, Beckett LA, Schenker MB. Residential proximity to naturally occurring asbestos and mesothelioma risk in California. American Journal of Respiratory and Critical Care Medicine. 2005;172(8):1019-1025
  52. 52. Baumann F, Buck BJ, Metcalf RV, Mclaurin BT, Merkler DJ, Carbone M. The presence of asbestos in the natural environment is likely related to mesothelioma in young individuals and women from southern Nevada. Journal of Thoracic Oncology. 2015;10(5):731-737
  53. 53. Baumann FCM. Environmental risk of mesothelioma in the United States: An emerging concern-epidemiological issues. Journal of Toxicology and Environmental Health, Part B. 2016;19(5-6):231-249
  54. 54. Roushdy-Hammady I, Siegel J, Emri S, Testa JR, Carbone M. Genetic-susceptibility factor and malignant mesothelioma in the Cappadocian region of Turkey. Lancet. 2001;357(9254):444-445
  55. 55. Oczypok EA, Sanchez MS, Van Orden DR, Berry GJ, Pourtabib K, Gunter ME, et al. Erionite-associated malignant pleural mesothelioma in Mexico. International Journal of Clinical and Experimental Pathology. 2016;9(5):5722-5732
  56. 56. Fraire AE, Lach-Pasko E, Greenberg D, Spjut HJ, Dodson RF, Williams G. Effect of erionite on the pleural mesothelium of the fischer 344 Rat. Laboratory Animal Investigation. 1997;111(5):1375-1380. DOI: 10.1378/chest.111.5.1375
  57. 57. Ozesmi M, Patiroglu TE, Hillerdal G, Ozesmi C. Peritoneal mesothelioma and malignant lymphoma in mice caused by fibrous zeolite. British Journal of Industrial Medicine. 1985;42(11):746-749
  58. 58. Hillegass JM, Miller JM, MacPherson MB, Westbom CM, Sayan M, Thompson JK, et al. Asbestos and erionite prime and activate the NLRP3 inflammasome that stimulates autocrine cytokine release in human mesothelial cells. Particle and Fibre Toxicology. 2013;10(1):1-14
  59. 59. Paoletti I, Batisti D, Bruno C, Di Paola M, Gianfagna AM, Mastrantonio M, et al. Unusually high incidence of malignant pleural mesothelioma in a town of eastern Sicily: An epidemiological and environmental study. Archives of Environmental Health. 2000;55(6):392-398
  60. 60. Suzuki YKN. Malignant mesothelioma induced by asbestos and zeolite in the mouse peritoneal cavity. Environmental Research. 1984;35(1):277-292. DOI: 10.1016/0013-9351(84)90136-1
  61. 61. Grosse Y, Loomis D, Guyton KZ, Lauby-Secretan B, El Ghissassi F, Bouvard V, et al. International Agency for Research on Cancer Monograph Working Group. Carcinogenicity of fluoro-edenite, silicon carbide fibres and whiskers, and carbon nanotubes. The Lancet Oncology. 2014;15(13):1427-1428
  62. 62. Cardile V, Renis M, Scifo C, Lombardo L, Gulino R, Mancari B. PA. Behaviour of the new asbestos amphibole fluoro-edenite in different lung cell systems. The International Journal of Biochemistry & Cell Biology. 2004;36(5):849-860. DOI: 10.1016/j.biocel.2003.09.007
  63. 63. Turci F, Tomatis M, Gazzano E, Riganti C, Martra G, Bosia A, et al. FB. Potential toxicity of nonregulated asbestiform minerals: Balangeroite from the western Alps. Part 2: Oxidant activity of the fibers. Journal of Toxicology and Environmental Health, Part A. 2005;68(1):21-39
  64. 64. Turci F, Tomatis M, Compagnoni R, Fubini B. Role of associated mineral fibres in chrysotile asbestos health effects: The case of balangeroite. The Annals of Occupational Hygiene. 2009;53(5):491-497
  65. 65. Piolatto G, Negri E, La Vecchia C, Pira E, Decarli A, Peto J. An update of cancer mortality among chrysotile asbestos miners in Balangero, northern Italy. British Journal of Industrial Medicine. 1990;47(12):810-814
  66. 66. IARC Working Group on the Evaluation of Carcinogenic Risks to. IARC Monographs on the Evaluation of Carcinogenic Risks to Humans. Lyon: IARC; 1988. pp. 39-153
  67. 67. Mast RW, Hesterberg TW, Glass LR, McConnell EE, Anderson R, Bernstein DM. Chronic inhalation and biopersistence of refractory ceramic fiber in rats and hamsters. Environmental Health Perspectives. 1994;102(Suppl. 5):207-209
  68. 68. Lacourt A, Gramond C, Audignon S, Ducamp S, Févotte J, Ilg AGS, et al. Pleural mesothelioma and occupational coexposure to asbestos, mineral wool, and silica. American Journal of Respiratory and Critical Care Medicine. 2013;187(9):977-982
  69. 69. Martini F, Corallini A, Balatti V, Sabbioni S, PC& TM. Simian virus 40 in humans. Infect Agents Cancer. 2007;2:13-25. DOI: 10.1186/1750-9378-2-13
  70. 70. Shah KNN. Human exposure to SV40: Review and comment. American Journal of Epidemiology. 1976;103(1):12
  71. 71. Rotondo JC, Mazzoni E, Bononi I, Tognon M, Martini F. Association between simian virus 40 and human tumors. Frontiers in Oncology. 2019;9(July):1-19
  72. 72. Cacciotti P, Libener R, Betta P, Martini F, Porta C, Procopio A, et al. SV40 replication in human mesothelial cells induces HGF/Met receptor activation: A model for viral-related carcinogenesis of human malignant mesothelioma. Proceedings of the National Academy of Sciences of the United States of America. 2001;98(21):12032-12037
  73. 73. Manfredi JJ, Dong J, Liu WJ, Resnick-Silverman L, Qiao R, Chahinian P, et al. Evidence against a role for SV40 in human mesothelioma. Cancer Research. 2005;65(7):2602-2609
  74. 74. Giuranno L, Ient J, De Ruysscher D, Vooijs MA. Radiation-induced lung injury (RILI). Frontiers in Oncology. 2019;9:1-16
  75. 75. Small GR, Nicolson M, Buchan K, Broadhurst P. Pericardial malignant mesothelioma: A latent complication of radiotherapy? European Journal of Cardio-Thoracic Surgery. 2008;33(4):745-747
  76. 76. Li X, Brownlee NA, Sporn TA, Mahar A, Roggli VL. Malignant (diffuse) mesothelioma in patients with hematologic malignancies: A clinicopathologic study of 45 cases. Archives of Pathology & Laboratory Medicine. 2015;139(9):1129-1136
  77. 77. Roggli VL, Oury TD. MEJ. Malignant mesothelioma in women. Anatomic Pathology. 1997;2:147-163
  78. 78. Horie A, Hiraoka K, Yamamoto O, Haratake J, Tsuchiya T. SH. An autopsy case of peritoneal malignant mesothelioma in a radiation technologist. Acta Pathologica Japonica. 1990;40(1):57-62
  79. 79. Borczuk AC, Pei J, Taub RN, Levy B, Nahum O, Chen J, et al. Genome-wide analysis of abdominal and pleural malignant mesothelioma with DNA arrays reveals both common and distinct regions of copy number alteration. Cancer Biology Therapy. 2016;17(3):328-335. DOI: 10.1080/15384047.2016.1145850
  80. 80. Kodama Y, Hoshi S, Minami M, Kiso M, Takezawa T, Arai T, et al. Malignant mesothelioma associated with chronic empyema with elevation of serum CYFRA19: A case report. Bioscience Trends. 2008;2(6):250-254
  81. 81. Livneh A, Langevitz P. PM. Pulmonary associations in familial Mediterranean fever. Current Opinion in Pulmonary Medicine. 1999;5(5):326-331
  82. 82. Carbone M, Yang H, Pass HI, Krausz T, Testa JR, Gaudino G. BAP1 and cancer chromosome microarray and genetic linkage analyses on the L and W families implicated. Nature Reviews Cancer. 2014;13(3):153-159
  83. 83. Jensen DE, Proctor M, Marquis ST, Gardner HP, Ha SI, Chodosh LA, et al. BAP1: A novel ubiquitin hydrolase which binds to the BRCA1 RING finger and enhances BRCA1-mediated cell growth suppression. Oncogene. 1998;16(9):1097-1112
  84. 84. Bononi A, Giorgi C, Patergnani S, Larson D, Verbruggen K, Tanji M, et al. BAP1 regulates IP3R3-mediated Ca2+ flux to mitochondria suppressing cell transformation. Nature. 2017;546(7659):549-553
  85. 85. Murali R, Wiesner T, Scolyer RA. Tumours associated with BAP1 mutations. Pathology. 2013;45(2):116-126
  86. 86. Kadariya Y, Cheung M, Xu J, Pei J, Sementino E, Menges CW, et al. Bap1 is a bona fide tumor suppressor: Genetic evidence from mouse models carrying heterozygous germline Bap1 mutations. Cancer Research. 2016;76(9):2836-2844
  87. 87. Testa JR, Cheung M, Pei J, Below JE, Tan Y, Sementino E, et al. Germline BAP1 mutations predispose to malignant mesothelioma. Nature Genetics. 2012;43(10):1022-1025
  88. 88. Ohar JA, Cheung M, Talarchek J, Howard SE, Howard TD, Hesdorffer M, et al. Germline BAP1 mutational landscape of asbestos-exposed malignant mesothelioma patients with family history of cancer. Cancer Research. 2016;76(2):206-215
  89. 89. Xu J, Kadariya Y, Cheung M, Pei J, Talarchek J, Peng H, et al. Germline mutation of Bap1 accelerates development of Asbestos-induced malignant mesothelioma. Cancer Research. 2014;74(16):4388-4397
  90. 90. Baumann F, Flores E, Napolitano A, Kanodia S, Taioli E, Pass H, et al. Mesothelioma patients with germline BAP1 mutations have 7-fold improved long-term survival. Carcinogenesis. 2015;36(1):76-81
  91. 91. Wiesner T, Obenauf AC, Murali R, Fried I, Griewank KG, Ulz P, et al. Germline mutations in BAP1 predispose to melanocytic tumors. Nature Genetics. 2011;43(10):1018-1022
  92. 92. Tsao H. Genotypic and phenotypic features of BAP1 cancer syndrome: A report of 8 new families and review of cases in the literature. JAMA Dermatology. 2017;153(10):999-1006
  93. 93. Farzin M, Toon CW, Clarkson A, Sioson L, Watson N, Andrici J, et al. Loss of expression of BAP1 predicts longer survival in mesothelioma. Pathology. 2015;47(4):302-307
  94. 94. Nasu M, Emi M, Pastorino S, Tanji M, Powers A, Luk H, et al. High incidence of somatic BAP1 alterations in sporadic malignant mesothelioma masaki. Journal of Thoracic Oncology. 2015;10(4):565-576
  95. 95. Klebe S, Driml J, Nasu M, Pastorino S, Zangiabadi A, Henderson D, et al. BAP1 hereditary cancer predisposition syndrome: A case report and review of literature. Biomarker Research. 2015;3(1):1-7
  96. 96. Rai K, Pilarski R, Cebulla CM, Cancer C, Science V, State TO, et al. Comprehensive review of BAP1 tumor predisposition syndrome. Clinical Genetics. 2016;89(3):285-294
  97. 97. Karczewski KJ, Francioli LC, Tiao G, Cummings BB, Alföldi J, Wang Q, et al. The mutational constraint spectrum quantified from variation in 141,456 humans. Nature. 2020;581(7809):434-443
  98. 98. Zhang N, Bai H, David KK, Dong J, Zheng Y, Cai J, et al. The Merlin/NF2 tumor suppressor functions through the YAP oncoprotein to regulate tissue homeostasis in mammals. Developmental Cell. 2010;19(1):27-38
  99. 99. Kanno S, Oda N, Abe M, Terai Y, Ito M, Shitara K, et al. Roles of two VEGF receptors, Flt-1 and KDR, in the signal transduction of VEGF effects in human vascular endothelial cells. Oncogene. 2000;19(17):2138-2146
  100. 100. Matullo G, Guarrera S, Betti M, Fiorito G, Ferrante D, Voglino F, et al. Genetic variants associated with increased risk of malignant pleural mesothelioma: A genome-wide association study. PLoS One. 2013;8(4):e61253
  101. 101. Cugliari G, Catalano C, Guarrera S, Allione A, Casalone E, Russo A, et al. DNA methylation of FKBP5 as predictor of overall survival in malignant pleural mesothelioma. Cancers (Basel). 2020;12(11):1-14
  102. 102. Petrilli AM, Fernández-Valle C. Role of Merlin/NF2 inactivation in tumor biology. Oncogene. 2016;35(5):537-548
  103. 103. Kane AB, Hurt RH, Gao H. The Asbestos-carbon nanotube analogy: An update. Toxicology and Applied Pharmacology. 2018;361:68-80
  104. 104. IARC. IARC Working Group on the Evaluation of Carcinogenic Risks to Humans. IARC Monographs on the Evaluation of Carcinogenic Risks to Humans. Ingested Nitrate and Nitrite, and Cyanobacterial Peptide Toxins. Vol. 94, IARC Monographs on the Evaluation of Carcinogenic Risks to Humans. Lyon: World Health Organization, International Agency for Research on Cancer; 2010
  105. 105. IARC. IARC Monographs on the Evaluation of Carcinogenic. Some Nanomaterials and Some Fibres. Vol. 111, IARC Monographs on the Evaluation of Carcinogenic Risks to Humans. Lyon: World Health Organization, International Agency for Research on Cancer; 2017
  106. 106. Numano T, Higuchi H, Alexander DB, Alexander WT, Abdelgied M, El-Gazzar AM, et al. MWCNT-7 administered to the lung by intratracheal instillation induces development of pleural mesothelioma in F344 rats. Cancer Science. 2019;110(8):2485-2492
  107. 107. Sakamoto Y, Nakae D, Fukumori N, Tayama K, Maekawa A, Imai K, et al. Induction of mesothelioma by a single intrascrotal administration of multi-wall carbon nanotube in intact male Fischer 344 rats. The Journal of Toxicological Sciences. 2009;34(1):65-76
  108. 108. Takagi A, Hirose A, Nishimura T, Fukumori N, Ogata A, Ohashi N, et al. Induction of mesothelioma in p53+/− mouse by intraperitoneal application of multi-wall carbon nanotube. The Journal of Toxicological Sciences. 2008;33(1):105-116
  109. 109. Muller J, Delos M, Panin N, Rabolli V, Huaux F, Lison D. Absence of carcinogenic response to multiwall carbon nanotubes in a 2-year bioassay in the peritoneal cavity of the rat. Toxicological Sciences. 2009;110(2):442-448
  110. 110. Donaldson K, Murphy FA, Duffin R, Poland CA. Asbestos, carbon nanotubes and the pleural mesothelium: A review of the hypothesis regarding the role of long fibre retention in the parietal pleura, inflammation and mesothelioma. Particle and Fibre Toxicology. 2010;7:1-17
  111. 111. Vance ME, Kuiken T, Vejerano EP, McGinnis SP, Hochella MF, Hull DR. Nanotechnology in the real world: Redeveloping the nanomaterial consumer products inventory. Beilstein Journal of Nanotechnology. 2015;6(1):1769-1780
  112. 112. Dowling A, Clift R, Grobert N, Hutton D, Oliver R, O’neill O, et al. Nanoscience and nanotechnologies: Opportunities and uncertainties. London The Royal Society & The Royal Academy of Engineering. 2004;46(July):618-618
  113. 113. Dahm MM, Schubauer-Berigan MK, Evans DE, Birch ME, Fernback JE, Deddens JA. Carbon nanotube and nanofiber exposure assessments: An analysis of 14 site visits. The Annals of Occupational Hygiene. 2015;59(6):705-723
  114. 114. Ma-Hock L, Strauss V, Treumann S, Küttler K, Wohlleben W, Hofmann T, et al. Comparative inhalation toxicity of multi-wall carbon nanotubes, graphene, graphite nanoplatelets and low surface carbon black. Particle and Fibre Toxicology. 2013;10(1):1-20
  115. 115. Jagirdar RM, Papazoglou ED, Pitaraki E, Kouliou OA, Rouka E, Giannakou L, et al. Cell and extracellular matrix interaction models in benign mesothelial and malignant pleural mesothelioma cells in 2D and 3D in-vitro. Clinical and Experimental Pharmacology & Physiology. 2021;48(4):543-552
  116. 116. Kuempel ED, Jaurand M, Morimoto Y, Pinkerton KE, Sargent LM, Vermeulen RCH, et al. Evaluating the mechanistic evidence and key data gaps in assessing the potential carcinogenicity of carbon nanotubes and nanofibers in humans. Critical Reviews in Toxicology. 2017;47(1):1-58

Written By

Evdoxia Gogou, Sotirios G. Zarogiannis, Dimitra Siachpazidou, Chryssi Hatzoglou and Konstantinos I. Gourgoulianis

Submitted: 21 January 2022 Reviewed: 11 February 2022 Published: 17 March 2022