Next Article in Journal
Multiple Myeloma: Possible Cure from the Sea
Next Article in Special Issue
Emergence of Resistance to MTI-101 Selects for a MET Genotype and Phenotype in EGFR Driven PC-9 and PTEN Deleted H446 Lung Cancer Cell Lines
Previous Article in Journal
Triplet Chemotherapy with Cisplatin versus Oxaliplatin in the CRITICS Trial: Treatment Compliance, Toxicity, Outcomes and Quality of Life in Patients with Resectable Gastric Cancer
Previous Article in Special Issue
Glioblastoma Microenvironment and Cellular Interactions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Implications of Gut Microbiota in Epithelial–Mesenchymal Transition and Cancer Progression: A Concise Review

1
College of Medicine, QU Health, Qatar University, Doha P.O. Box 2713, Qatar
2
Biomedical Research Center, Qatar University, Doha P.O. Box 2713, Qatar
*
Authors to whom correspondence should be addressed.
Cancers 2022, 14(12), 2964; https://doi.org/10.3390/cancers14122964
Submission received: 16 May 2022 / Revised: 7 June 2022 / Accepted: 8 June 2022 / Published: 16 June 2022
(This article belongs to the Special Issue EMT in Cancer)

Abstract

:

Simple Summary

Recently, the interactions between microbiota and the host have been reported to induce the onset and progression of human cancer via epithelial–mesenchymal transition (EMT). In contrast, some microorganisms can protect against cancer growth, indicating an anticancer therapeutic action of such microbiota. In the review, we summarize findings from the literature, exploring the underlying mechanisms by which pathogenic microorganisms induce EMT. We also highlight the potential of exploiting these complex interactions for developing new biological therapies.

Abstract

Advancement in the development of molecular sequencing platforms has identified infectious bacteria or viruses that trigger the dysregulation of a set of genes inducing the epithelial–mesenchymal transition (EMT) event. EMT is essential for embryogenesis, wound repair, and organ development; meanwhile, during carcinogenesis, initiation of the EMT can promote cancer progression and metastasis. Recent studies have reported that interactions between the host and dysbiotic microbiota in different tissues and organs, such as the oral and nasal cavities, esophagus, stomach, gut, skin, and the reproductive tract, may provoke EMT. On the other hand, it is revealed that certain microorganisms display a protective role against cancer growth, indicative of possible therapeutic function. In this review, we summarize recent findings elucidating the underlying mechanisms of pathogenic microorganisms, especially the microbiota, in eliciting crucial regulator genes that induce EMT. Such an approach may help explain cancer progression and pave the way for developing novel preventive and therapeutic strategies.

1. Introduction

Epithelial cells are apicobasal polarized cells that function as physical barriers. They are tightly bound to adjacent cells, and the extracellular matrix (ECM) is regulated by E-cadherins and cytokeratins, respectively [1,2]. However, under certain conditions, including developmental processes, wound healing, repair, and tumor progression, epithelial cells lose their high degree of plasticity and attain migratory and invasive capabilities [3]. During the alteration of epithelial cells, junctional proteins are relocalized, or a more severe event occurs, such as the epithelial–mesenchymal transition (EMT) initiation [4,5]. During EMT, epithelial cells undergo loss of cell-to-cell junction and reorganization of the actin cytoskeleton; thus, nonmotile epithelial cells are converted to motile and invasive mesenchymal phenotypic cells [6]. Morphologically, epithelial cells lose their polygonal phenotype and acquire an elongated fibroblast morphology; these events are regulated by vimentin, fibronectin, and N-cadherin [4,5]. EMT is characterized by loss of E-cadherin and translocation of β-catenin from the cell membrane to the nucleus, followed by activation of several mesenchymal markers (e.g., vimentin, fibronectin, and N-cadherin) [4,5].
Though several growth factors activate signaling pathways to control EMT gene expression, some EMT-signaling pathways are regulated by microbial pathogens [7,8].
Previous studies reported that microbe invasion might alter the transforming growth factor β (TGFβ); thus, the TGFβ receptor phosphorylates and activates transcription factors Smad-2 and Smad-3, which heterodimerize with Smad-4 to form the Smad complex [9,10,11]. The Smad complex recruits the Ras-MAPK pathway leading to cell growth; proliferation; differentiation; migration; and, therefore, cancer progression [12,13].
In this review, we present a brief overview of the human gut microbiome, focusing on gut dysbiosis during EMT. We present data from the literature that shed light on their possible role in this crucial event, further triggering carcinogenesis and its progression.

2. Microbiota

Of the total human cell count, around 90% are associated with the presence of microbiota, while the remaining 10% are microbiome-free [14]. Nevertheless, it is postulated that the number of microbial genes is approximately ten times higher than the number of human genes [14]. Primarily located in the gut, the microbes play a vital role in nutrient uptake [15] and influence the development of healthy intestinal immune responses [16]. Any modification or change in the microbiota composition disrupts the microbe–immune system relationship, further inducing the onset and development of several human inflammatory disorders that may lead to EMT [17,18].
The Human Microbiome Project (HMP) was a two-phase research initiative that used metagenomics and whole-genome sequencing in the first phase to recognize and distinguish the whole human microbiota [19]. In the second phase, the project revealed the role of microbes in human diseases using multiple omics techniques [19,20]. Although alterations in genes regulating DNA repair are mainly responsible for the onset and progression of tumorigenesis, the HMP indicated a role of dysbiotic microbiota in cancer progression [19]. With recent advancements, the use of genomics, epigenomics, proteomics, metabolomics, and transcriptomics elucidated host–microbiota interactions and their underlying mechanisms in human diseases; however, its role in carcinogenesis is still nascent. While viruses express active oncoproteins that can induce cell transformation leading to tumor formation or progression, dysbiosis-induced carcinogenesis arises after multiple hits [21]. An in vivo study using gnotobiotic (including germ-free) mouse models reported that microbes affect metabolism and inflammation, provoking the onset and progression of cancer [22].
Due to the extensive presence of microbes in the gut, several studies have primarily focused on the effects of altered microbiota in colorectal cancer pathogenesis [23,24,25,26,27,28,29,30,31,32]. Nevertheless, recent investigations have shown a correlation between dysbiosis and other cancers, including breast, oral, lung, skin, and reproductive tract [33,34,35,36,37,38,39,40,41]. The following section discusses the interplay between the host and the microbiota in triggering the onset of cancer via EMT.

3. Microbiota-Induced Epithelial–Mesenchymal Transition

Microbes induce EMT by attaching to the mucosal layers and trigger the breakdown of intercellular adhesion between epithelial cells. Bacterial adhesins bind to epithelial proteins’ E-cadherin/catenin complex, thus altering cell polarity and downstream signaling pathways, leading to EMT [42]. A study by Chen and colleagues reported that immunosuppression due to severe inflammation that overwhelms both regulatory T-cells and dendritic cells was significantly associated with the onset of EMT [43,44]. In the colon, Fusobacterium nucleatum (F. nucleatum) enhances the release of inflammatory cytokines [45]; in the urogenital tract, infection with Lactobacillus spp. triggers the release of interleukins [39].
One of the most common bacteria the Gram-negative, microaerophilic bacteria, Helicobacter pylori (H. pylori) is present in the digestive tract in approximately 50% of the population worldwide [46,47]. In addition to its causative role in inflammation and ulceration in gastric epithelial cells, H. pylori can trigger toll-like receptors-2 and -5 (TLR2 and TLR5) to activate NFκB [48]. On the other hand, the virulent cytotoxic factors of H. pylori, CagA, and VacA can disrupt epithelial cell function. CagA disrupts the apical junctional complex and actin-cytoskeletal rearrangements. In contrast, VacA destroys the barrier function of tight junctions, leading to loss of epithelial cell-to-cell adhesion and loss of cell polarity [49,50,51]. Brandt and colleagues [52] reported that CagA could induce the release of IL-8 via the Ras/Mek/Erk/NFκB signaling pathways (Figure 1). Following this, another study by Yin et al. [53] showed that pathogenic strains of H. pylori enhance the expression of vimentin, Snail, and Slug supported by upregulated levels of gastrin; MMP7; and soluble heparin-binding epidermal growth factor. The studies support the role of H. pylori in the remodeling of actin filaments leading to the onset of EMT [52,53].
Moreover, E. coli is present in the gastrointestinal tract within a few hours after birth and generally harmonizes with its human hosts [54]. However, during the loss of intestinal barrier permeability due to the relocalization of junctional proteins, E. coli triggers the onset of diarrhea [55]. In chronic cases, the event can lead to EMT. Studies reported that diffusely adherent E. coli (DAEC) could infect intestinal epithelial cells and promote EMT by activating MAPK and PI3K pathways (Figure 1) [56,57,58]. In addition, the bacteria will stimulate the overexpression of HIF-1α protein, accentuating loss of E-cadherin and cytokeratin 18 and upregulation of fibronectin, signifying a possible role of E. coli in EMT [59].
In contrast, several bacterial products, such as lipopolysaccharide (LPS), flagellin, and muramyl dipeptides (MDP), are extensively studied. LPS, a vital part of the outer membrane of Gram-negative bacteria, is an endotoxin that binds to TLR4 [60]. Although earlier studies reported that LPS-induced EMT is very scarce, Zhao et al. demonstrated that LPS reduced the expression of the epithelial biomarker E-cadherin in intrahepatic biliary epithelial cells [61]. In contrast, the expression of mesenchymal markers S100A4 and α-SMA was enhanced [61]. More importantly, this investigation reported that LPS leads to overexpression of TGFβ-1 [61], an important inducer of EMT via Smad 2/3 [62]. Silencing of Smad 2/3 expression in these cells triggered E-cadherin expression and inhibited S100A and α-SMA deregulation, indicating that LPS induced EMT via the TGFβ1/Smad2/3 pathway [61]. Similar to LPS, flagellin and MDP are also found to trigger EMT. Both flagellin and MDP trigger the NFκB and MAPK signaling pathways [63,64]. In addition, flagellin stimulates the production of TGF-β and TGFβ1, which are known inducers of EMT [65]. Similarly, MDP also induces the expression patterns of genes responsible for invasive cell growth in intestinal epithelial cells and EMT [66].

4. Microbiota-Enhanced Carcinogenesis via Epithelial–Mesenchymal Transition

Dysbiosis is associated with host inflammatory responses and EMT in various sites favoring cancer progression. In cancer cells, EMT activation is related to the presence of altered infiltrating tumor-associated macrophages (TAMs), which produce soluble growth factors and inflammatory cytokines and promote extracellular matrix remodeling, angiogenesis, immunosuppression, and cancer cell invasion [67]. In addition, several studies reported that cancer microbiota initiates EMT and tumorigenesis via metabolic reprogramming (Figure 1) [68,69,70]. In the following subsections, we will focus on the presence of microbial colonies in different anatomical sites and their underlying signaling mechanisms responsible for triggering EMT leading to cancer progression.

4.1. Respiratory Tract Microbiota

The nasal epithelium is predominated by bacteroidetes, firmicutes, proteobacteria, and actinobacteria [71]. However, a wide variation in the microbial composition has been reported based on several factors, including humid environment, temperature, and localization within the respiratory tract [72]. The nasal microbiota can alter the expression and functions of regulators of the olfactory signaling transduction pathways [71], in addition to the onset of allergic rhinitis and chronic rhinosinusitis [72]. Microbes act as epithelial barriers in the nasal cavity and can promote tissue-remodeling [73]. During microbial infections, the mucociliary clearance is altered, and nasal microbiota is not removed from the airways; thus, they attach to the mucosal surface, form colonies, and produce soluble virulence-associated factors [74]. Ziesemer et al. reported that alpha-hemolysin, a cytotoxic agent released by Staphylococcus aureus (S. aureus) in human airway epithelial cells, enhanced actin filament remodeling due to disruption of cell-to-cell contact and the focal adhesions leading to the augmented penetrability of the epithelial layer [75]. Moreover, S. aureus is involved in nasal polyposis pathogenesis [76]; nasal polyps lack expression of E-cadherin and occluding, while TGFβ and vimentin are overexpressed compared with healthy nasal mucosa [77], indicating a role of S. aureus in EMT.
On the other hand, the lung is primarily composed of Bacteroidetes and Firmicutes [78,79]. During respiratory diseases, mucus production presents suitable environmental and nutrient conditions for the microbes to thrive; hence, the microbial composition is altered [80]. This altered microbial composition promotes genotoxic and virulent effects, leading to deregulated metabolism, inflammation, and immune response, features of lung cancer development [80]. A recent study by Jin and colleagues [81] used lung adenocarcinoma mouse models with Kras mutations and p53 deletion to study microbiota-induced inflammation in different myeloid cells. The study reported that in adenoviral (Sftpc-Cre) infected mice, the local microbiota activated myeloid cells (neutrophils) to enhance the production of IL-1β, IL-23, and γδ T cells to stimulate inflammation and tumor cell proliferation through IL-17. Thus, germ-free or antibiotic-exposed mice are drastically protected against lung adenocarcinoma compared to adenoviral (Sftpc-Cre) infected mice [81].
Previous studies have investigated the role of gut microbiota in extra-gastrointestinal tumors [33,35], including lung cancer [34,36,37]. Recently, Enterococcus and Bifidobacterium were associated with the onset of lung cancer and, therefore, suggested as a potential diagnostic biomarker in lung cancer [82]. On the other hand, differential expression of gut microbiota was also observed in lung cancer; the expressions of Escherichia-Shigella, Enterobacter, Dialister, Kluyvera, and Faecalibacterium were reduced in lung cancer patients, while Veillonella, Fusobacterium, and Bacteroides were augmented in comparison with healthy individuals [37]. Moreover, non-small-cell lung cancer (NSCLC) patients had higher levels of gut bacteria when compared with healthy controls [83]. However, on the contrary, downregulated levels of gut butyrate-producing bacteria (Clostridium leptum, Faecalibacterium prausnitzii, Ruminococcus, and Clostridial cluster I spp.) were recently reported in NSCLC patients [84]. Liu et al. [85] carried out 16S ribosomal RNA (rRNA) gene amplicon sequencing in 30 lung cancer patients compared with 16 healthy individuals. They reported that gut microbiota dysbiosis in lung cancer correlates with altered metabolic and immunologic functions involved in the development and progression of lung cancer.
Similarly, a recent study by Zheng and colleagues utilized the 16S rRNA gene sequencing analysis and revealed the microbiota spectrum of lung cancer patients [86]. The study further reported a potential gut microbial signature for the prediction of early-stage lung cancer [86]. Another recent investigation demonstrated that prebiotics and probiotics have a latent protective effect on lung carcinogenesis [87]. Although studies have reported altered gut microbiome as a potential diagnostic and prognostic marker [88], further studies are warranted to examine the underlying mechanisms of the gut microbiome in the onset and progression of lung cancer.
Nonetheless, studies have also indicated an interaction between the gastrointestinal (GI) and respiratory tracts known as the gut–lung axis by altering microbial and immune functions [89] through a complex bidirectional axis involving blood and lymphatic circulation [90,91]. Dysregulation in the gut–lung axis is implicated in pathogen colonization, tissue damage, and the onset of carcinogenesis [92,93]. There are different pathways involved in the role of the gut–lung axis in lung cancer pathogenesis. TLRs on the intestinal epithelial cells surface identify microbial ligands and induce TLR innate-adaptive immunity; immune cell migration triggers the gut mucous membranes [94]. Inflammation is another mechanism involved in gut–lung-axis-induced lung cancer; microbes migrate from the GI tract to the bloodstream via the mucosal barrier and induce lung inflammation, further augmenting the innate systemic response [95,96,97]. In addition, secondary metabolites produced from bile acids by gut bacteria and alteration of the gut microbiota cause DNA damage, produce toxins, and initiate cancer development; deregulated metabolism triggers toxic metabolite formation in the lungs and contributes to the development of lung cancer [98,99,100]. With the potential role of the gut–lung axis in lung cancer pathogenesis, the possibility of its manipulation for developing biological therapeutic agents needs to be studied further.

4.2. Gastrointestinal (GI) Tract Microbiota

Recently, it has been reported that microbial pathogens, especially intestinal microorganisms, play an essential role in carcinogenesis; intestinal dysbiosis can induce immune response triggering chronic inflammation and, in adverse conditions, leading to cancer progression [101].
Oral cancer arises from the oral mucosa, and approximately 15% of the cases are attributed to oral microbial dysbiosis [102]. The oral cavity is inhabited by various microbial species, including Porphyromonas gingivalis (P. gingivalis), F. nucleatum, Streptococci, Peptostreptococci, and Prevotella [103]. Dysbiosis of the oral microbiome alters the immune response resulting in an increased risk of the onset of periodontal diseases and oral squamous cell carcinoma (OSCC) [104,105,106]. Chronic infection in oral cells by P. gingivalis induces the expression of CD44 and CD133, which activate matrixins (MMPs-1 and -10) along with Slug, Snail, and Zeb1 leading to EMT [107,108]. This process of chronic-infection-induced EMT in the oral cavity results in oral cells developing invasive and migrative properties [107,108].
The role of microbiota has been assessed in esophageal cancer; in comparison to normal esophagus tissue, reduced microbial diversity is reported in Barrett’s esophagus, esophageal adenocarcinoma (EAC), and esophageal squamous cell carcinoma (ESCC) [109,110,111,112,113,114,115,116,117,118]. In EAC, Akkermansia and Gram-negative bacteria, Lactobacilli, Prevotella, Leptotrichia, and Enterobacteriaceae are augmented with loss of Streptococci [109,114,119]; in ESCC, Streptococci, Fusobacteria, Veillonella, and P. gingivalis are abundant with reduced Lautropia, Bulleidia, Catonella, Corynebacterium, Moryella, Peptococcus, Treponema, and Cardiobacterium [111,115,116,120]. An in vivo study using a xenograft model reported microbial pathogens to play a role in increased uptake of metabolic glucose in addition to EMT in the esophagus [121]. Moreover, dietary intake is reported to affect the microbial composition in esophageal cancer pathogenesis. Kaakoush and colleagues performed an in vivo study using Sprague Dawley rats; the rats were given an obesogenic diet and had an altered esophageal microbiota associated with chronic gastrointestinal diseases compared with normal diet-fed rats [122]. Another recent in vivo study involved transgenic mice (L2-IL1B mice) fed a high-fat diet; the study reported dysbiosis of the esophageal and gut microbiota resulting in inflammation and development of esophageal tumors in comparison with mice fed a normal diet [123]. Moreover, Riboflavin, a vitamin B2 supplement, impacts the balance between gut microbiota and esophageal mucosal integrity [124]. In vivo studies reported that riboflavin deficiency alters the gut microbiota and leads to esophageal epithelial atrophy [125,126]. The role of H. pylori in esophageal cancer is conflicting. Although the reduced H. pylori incidence is associated with an increased risk of EAC, there was no significant association between H. pylori infection and ESCC [127,128,129]. However, one study reported that H. pylori infection is associated with ESCC in the non-Asian population; in the Asian population, it showed a converse relation [128]. On the contrary, studies in the US and Swedish populations failed to find an association between H. pylori infection and EAC incidence [130,131]. These studies indicate a need to investigate the role of H. pylori in the onset and development of esophageal cancer.
However, H. pylori is the most common cause of gastric cancer and is classified as a class I carcinogen involved in the onset of gastric cancer pathogenesis by inducing inflammation and alteration of the gastric mucosal integrity [132,133,134,135,136]. Human gastric microbiota profiling revealed differential microbiota profiles between chronic gastritis, metaplasia, and gastric cancer, indicating that dysbiosis is associated with cancer progression [132,137,138,139,140,141,142,143]. Gastric colonization with H. pylori and Clostridium, Lactobacillus, and Bacteroides enhance inflammation with upregulated IL-11 expression and oncogenic genes, Ptger4, and Tgf-β, plausibly regulated by the yes-associated protein 1 (YAP1) [123,144]. Gastric microbiota analysis using 16S rRNA gene profiling demonstrated a distinct dysbiotic microbial community with plausible genotoxicity in gastric cancer in comparison to chronic gastritis [132]. In addition to the abundance of H. pylori in gastric cancer, oral-associated bacteria have also been found in patients with gastric cancer [136]. Several other investigations also reported the loss of H. pylori in gastric cancer in lieu of the dominant presence of Clostridium, Enterococci, Fusobacterium, Veillonella, Leptotrichia, Staphylococci, and Lactobacillus species [141,145]. Similarly, another study reported the presence of F. nucleatum to correlate with an overall worse prognosis in Laurens’s diffuse-type gastric cancer [146]. Lately, in vivo studies revealed that a high-fat diet in mice stimulates gastric dysbiosis and the enhanced presence of Lactobacilli, intestinal metaplasia, STAT3, and accumulation of β-catenin; these changes provide a protumorigenic gastric microenvironment leading to the onset and development of gastric cancer [147,148].
Contrary to gastric cancer, the role of H. pylori in colorectal cancer (CRC) pathogenesis is unclear. Dysbiosis of the gut microbiota is reported in tissues of CRC when compared with normal tissue [23,24,25,26,27,28,29,30,31,32]. Sears and Pardoll suggested the “alpha-bug” (enterotoxigenic Bacteroides fragilis) hypothesis for colorectal cancer, where they found that oncogenic microorganisms can modify the mucosal immune response and colonic bacterial community to promote colorectal cancer [149]. On the other hand, Tjalsma and colleagues [150] proposed another model for colorectal cancer, known as the “driver passenger”, where they explained that tumors induced by microbes (driver) are subsequently replaced by other symbiotic microbes (passengers) and can alter the local infectious environment, further promoting tumorigenesis. Different gut microbiota species are found in different stages of CRC progression; Gram-positive bacteria (Firmicutes and Actinobacteriaphyla) are dominant in premalignant adenomas, whereas Gram-negative bacteria (Enterobacteriaceae, Proteobacteria, Burkholderiales, and Sutterellai) are dominant in CRC [151]. Furthermore, the microbe Oscillospira is lost during the transition from advanced adenoma to early CRC [30]. One commonly detected microbial pathogen in CRC is F. nucleatum, which correlates with an elevated risk of CRC recurrence and chemoresistance [45,152,153]. F. nucleatum adheres to the colonic mucosa and interacts with Fap2 and integrin α21 promoting cell proliferation and triggering the NF-κB pathway (Figure 1), in addition to the inhibition of natural killer cell response and accumulation of myeloid cells. These events alter the tumor microenvironment leading to microbial metastatic spread [45,154]. Furthermore, E. coli is also associated with the development of colon cancer; E. coli regulates the production of colibactin, a genotoxic E. coli strain resulting in DNA double-strand breaks, gut microbiota dysbiosis, stimulation of the NF-κB and Wnt/β-catenin pathways, as well as inflammation of the colonic mucosa, further stimulating cell proliferation [155,156]. In addition to these microbes, Fragilysin is another microbe present in the gut [157]. Fragilysin attaches to the epithelial receptors of the colon and initiates the NF-κB pathway, thus leading to an increase in colon cell growth, proliferation, and DNA damage [158,159]. On the other hand, Fragilysin also triggers cell proliferation and c-MYC activation by deregulating the Wnt/β-catenin signaling pathway via E-cadherin cleavage [158,159,160].
Likewise, a recent report indicates that oral and gut microbiota dysbiosis enhance bacterial invasion, which correlates with pancreatic cancer incidence [161]; however, studies are scarce on this particular topic. At the same time, other investigations reported that P. gingivalis in the oral cavity increased the risk of the onset of pancreatic ductal adenocarcinoma and cancer [162,163]. However, the role of H. pylori is contradictory in pancreatic cancer; while Wei et al. [164] suggested H. pylori as a risk factor for the development of pancreatic cancer, another study failed to detect H. pylori in pancreatic tissue or fluid by PCR [165]. It is also evidenced that H. pylori secretes cytotoxins and vacuolins, and induces chronic inflammation and DNA damage, leading to pancreatic carcinogenesis [165,166]. Furthermore, 16s rRNA gene sequencing in pancreatic ductal adenocarcinoma identified 13 different microbe phyla, of which the most abundant were Proteobacteria, followed by Bacteroids and Firmicutes [167]. Commonly, duodenal or biliary bacterial reflux promotes translocation and colonization of the gut microbiota in the pancreas [167], enhancing the development and progression of pancreatic cancer [168].

4.3. Female Reproductive Tract Microbiota

The cervicovaginal tract comprises a diversified and complex microbial community named cervicovaginal microbiome (CVM), regulating different physiological disorders [169,170]. Although the CVM is composed of different microbe communities, it is highly dominated by the genus Lactobacillus (Lactobacillus crispatus, Lactobacillus iners, Lactobacillus gasseri, or Lactobacillus jensenii) [171,172]. In addition to maintaining tissue homeostasis [173] and a local pH lesser than 4.5 [174], lactobacilli adhere to epithelial cells by forming microcolonies and serve as a barrier to protect the genital environment from infectious pathogens [175], counteracting bacterial vaginosis, yeast infections, and sexually transmitted diseases (STDs) [176,177]. The imbalance of the CVM triggers abnormal cell proliferation, chronic inflammation, genome instability, STDs, premature births, and cancers of the vaginal tract [40,178,179,180]. Enhanced vaginal dysbiosis induces proinflammatory cytokines and chemokines production, followed by an inflammatory response [181] and dysregulation of the immune response favoring a tumor-promoting microenvironment [182,183]. The presence of Atopobium vaginae and Porphyromonas sp. in the reproductive tract, along with an increased vaginal pH (>4.5), correlated with the onset of endometrial cancer [184]. On the other hand, in cervical cancer, Laniewski et al. [39] reported that a low abundance of lactobacilli is associated with increased vaginal pH and enhanced secretion of various inflammatory cytokines, including interleukins (IL-2, IL-4, and IL-36γ), MIP-1β, IP-10, Flt-3L, and sCD40L. A study by Mitra and colleagues [185] reported high bacterial variation and loss of lactobacilli to be associated with cervical intraepithelial neoplasia (CIN) progression and cytological lesion severity. In addition, several studies have linked vaginal dysbiosis with human papillomavirus (HPV) infection in different grades of CIN and cervical cancer [185,186,187,188]. A study by Kwasniewski [189] reported dysbiosis of vaginal microbiota to induce the development of HPV-induced cervical cancer, indicating a role of vaginal microbiota in regulating viral persistence. Other studies also reported an association of reduced lactobacilli with an increased risk of HPV infection and bacterial vaginosis [186,190]. Lactobacilli reduce microbiome composition, triggering inflammation that can stimulate the expression of high-risk HPV oncogenes (E6 and E7) and malignant cell proliferation [191]. Studies also reported differential expression of microbiota in ovarian cancer tissues compared with normal tissues. Chronic infection with Proteobacteria and Firmicutes induces an inflammatory immune response leading to the onset and progression of ovarian carcinogenesis [38,41].
Since the gut microbiota shares approximately 30% of bacterial species, including Firmicutes, Bacteroidetes, Proteobacteria, Actinobacteria, and Fusobacteria [191], and regulates circulating estrogen (estrobolome), it is suggested that there is a crosslink between the gut/estrobolome and related risk of vaginal diseases including malignancies [191,192,193,194]. The gut microbiome may be regarded as a reservoir for vaginal microbes. The Group B streptococcus is present in the gut; however, if present in the vagina of pregnant women, it can induce premature delivery [195]. Enhanced levels of lactobacilli in the vagina reduce bacterial vaginosis [196]. However, the intake of oral probiotics was found to inhibit bacterial vaginosis, indicating an influence on the gut microbiome in the vagina [197]. In addition, Lactobacillus, Bacteroides, Bifidobacterium, and Akkermansia are associated with enhanced levels of short-chain fatty acids (SCFA) [198]; a differential role of SCFA has been shown between the gut and the vagina [191]. In the gut, SCFAs have anti-inflammatory characteristics and regulate the intestinal epithelial barrier [198]; whereas, in the vagina, SCFAs’ expression might be linked with several proinflammatory biomarkers [199]. In ovarian cancer, Xu et al. [200] demonstrated that intestinal dysbiosis activates tumor-associated macrophages and increases circulating levels of proinflammatory cytokines (IL-6 and TNF-α), promoting the onset of EMT.
Nevertheless, vaginal pathogens inducing diseases of the gut are still inconclusive. While studies reported delivery via the vagina or cesarean section to protect against asthma and gastroenteritis [201,202], another study did not find any association between the mode of delivery and respiratory or gut diseases [203]. While the gut microbiome is contemplated as one of the vital regulators of circulating estrogens, studies supporting the role of estrogen-related signaling and high-risk HPV-induced cancer are nascent and warrant further research [192,193,194]. Table 1 summarizes the roles of various gut microbiota in the onset of some common cancers.

5. Microbiome-Based Therapies (Biotherapy)

Gut microbiota in EMT-induced carcinogenesis is also involved in response to cancer therapy and toxicities [219,220]. The gut microbiota dysbiosis can modify both the systemic immune system and the response to chemotherapeutic agents [221,222]. However, cancer therapeutic drugs and antibiotics administration during the surgical or chemotherapeutic intervention can alter the gut microbiota. Moreover, chemotherapy and radiotherapy induce significant gut dysbiosis by destroying intestinal or colonic mucosa and altering several metabolic pathways leading to the risk of colitis [223,224,225].
To overcome these challenges, studies have focused on restoring the gut microbiota and helped pave the way for therapeutic strategies. For instance, fecal microbiota transplantation (FMT) was primarily used to treat Clostridioides difficile infection (CDI) by retention enemas and became common practice over the last decade [226,227]. FMT is administered through several ways, including infusion via nasogastric tube, oral capsules colonoscopy, and enema [228]; similar response rates were achieved for both oral administration and colonoscopy [229]. FMT is emerging as a candidate therapeutic option for treating several gut dysbiotic nonmalignant diseases, including irritable bowel syndrome, inflammatory bowel disease, multidrug-resistant diseases, metabolic syndrome, diabetes, nonalcoholic fatty liver disease, neuropsychiatric disorders, and autoimmune diseases [230,231,232,233]. However, although there are clinical trials observing the use of FMT against cancer in clinical practice, this still lies nascent [234,235,236].
On the other hand, probiotics involve the intake of bacteria or a combination of live organisms via supplements to maintain the normal microflora in the body [237]. Research has explored several commercially available probiotics in clinical trials, especially in CRC tumorigenesis. Interestingly, such studies demonstrated the efficacy of probiotic VSL#3 in CRC [238,239]; contrarily, another investigation reported that VSL#3 alters the mucosal microbial composition and enhances tumor growth [240]. In addition, although the effect of probiotic administration has been examined in several clinical trials in cancer patients, the studies majorly focused on the analysis of microbe dysbiosis [241,242,243,244,245]. Hence, more studies are required to assess the differential outcome of probiotics against cancer.
As previously stated, diet plays a role in gut microbiota composition and their metabolomic and transcriptomic profiles [147,246,247]. Numerous reports have indicated diet intake as a potential anticancer intervention [246,248,249]. On the other hand, prebiotics and postbiotics can also alter gut microbiota. Substances, including fructans, induce the growth of certain bacteria and modify SCFA levels within the gut; fructans were found to increase the efficacy of chemo- and radiotherapeutic agents in murine models [250]. In humans, use of postbiotics was studied against CRC and it was found that intake of butyrogenesis from high-fat-diet foods suppressed CRC carcinogenesis [251].
Finally, it is known that the use of antibiotics is associated with significant alteration in gut microbiota and worse clinical outcomes [252]. For example, patients with NSCLC demonstrated poor prognoses when given antibiotics before and after the start of treatment with immune checkpoint blockade [253]. Similarly, when administered anti-Gram-positive antibiotics, chronic lymphocytic leukemia patients had poor overall survival and response rates and earlier disease progression [253]. However, it might be useful to develop targeted antibiotics and bacteriophages to target the microbiota efficiently and improve therapeutic response selectively. In contrast, bacteriophages are the most significant and distinct members of the gut virobiota and have demonstrated efficiency in structuring the gut microbiota and targeting specific bacterial colonies [254,255]. Although these studies highlight the critical role of gut microbiota and biotherapy in the management of certain diseases, including cancer, additional studies are warranted to understand the underlying mechanisms and their plausible impact on the normal flora and immune system.

6. Conclusions

This review presents a concise outlook on the role of dysbiotic microbiota in EMT by altering transcription factors and deregulating signaling pathways, mainly STAT3, Wnt/β-catenin, and NF-κB. Although the role of microbes is well-defined in health and disease, their function in enhancing cancer progression via EMT is still nascent. Microbes inducing fibrin production or cancer have been implicated in EMT. Hence, understanding and unraveling the impact of the microbiota in inducing EMT and, therefore, cancer progression can help develop novel therapeutic regimens and biotherapies for human diseases, including cancers.

Author Contributions

Conceptualization, A.-E.A.M.; writing-original draft preparation, I.G.; writing-review and editing, S.V., S.P. and A.-E.A.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors would like to thank A. Kassab for her critical reading of the manuscript.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

Abbreviations

α-SMAAlpha smooth muscle actin
β-cateninBeta-catenin
γδ T cellsGamma delta T cells
CagACytotoxin-associated gene A
CINCervical intraepithelial neoplasia
c-MYCCellular myelocytomatosis
CRCColorectal cancer
CVMCervicovaginal microbiome
DAECDiffusely adherent
EACEsophageal adenocarcinoma
E-cadherinEpithelial cadherin
ECMExtracellular matrix
E. coliEscherichia coli
EMTEpithelial mesenchymal transition
ERKExtracellular-signal-regulated kinase
ESCCEsophageal squamous cell carcinoma
F. nucleatumFusobacterium nucleatum
Flt-3LFMS-like tyrosine kinase 3 ligand
GIGastrointestinal
HIF-1αHypoxia-inducible factor 1-alpha
H. pyloriHelicobacter pylori
HMPHuman Microbiome Project
HPVHuman papillomavirus
ILInterleukin
IP-10Interferon gamma-induced protein 10
KrasKirsten rat sarcoma viral oncogene homolog
LPSLipopolysaccharide
MAPKMitogen-activated protein kinase
MDPMuramyl dipeptides
MIPMacrophage Inflammatory Proteins
MMPsMatrix metalloproteases
N-cadherinNeural cadherin
NF-κBNuclear factor kappa light chain enhancer of activated B cells
NSCLCNon-small-cell lung cancer
OSCCOral squamous cell carcinoma
PCRPolymerase chain reaction
P. gingivalisPorphyromonas gingivalis
p53Tumor protein 53
PI3KPhosphatidylinositol 3-kinase
Ptger4Prostaglandin E Receptor 4
RasRat sarcoma virus
ROSReactive oxygen species
rRNARibosomal RNA
S100AS100 Calcium Binding Protein A1
S. aureusStaphylococcus aureus
SCFAShort chain fatty acids
SMADSuppressor of Mothers against Decapentaplegic
STATSignal transducer and activator of transcription
STDsSexually transmitted diseases
TAMsTumor-associated macrophages
TGFβTransforming growth factor β
TILsTumor-infiltrating lymphocytes
TLRToll-like receptor
TNFTumor necrosis factor
VacAVacuolating toxin A
WntWingless
YAP1Yes-associated protein 1
Zeb1Zinc Finger E-Box Binding Homeobox 1.

References

  1. Farquhar, M.G.; Palade, G.E. Junctional complexes in various epithelia. J. Cell Biol. 1963, 17, 375–412. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Lee, J.L.; Streuli, C.H. Integrins and epithelial cell polarity. J. Cell Sci. 2014, 127, 3217–3225. [Google Scholar] [CrossRef] [Green Version]
  3. Grünert, S.; Jechlinger, M.; Beug, H. Diverse cellular and molecular mechanisms contribute to epithelial plasticity and metastasis. Nat. Rev. Mol. Cell Biol. 2003, 4, 657–665. [Google Scholar] [CrossRef]
  4. Nieto, M.A. The ins and outs of the epithelial to mesenchymal transition in health and disease. Annu. Rev. Cell Dev. Biol. 2011, 27, 347–376. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Thiery, J.P.; Acloque, H.; Huang, R.Y.; Nieto, M.A. Epithelial-mesenchymal transitions in development and disease. Cell 2009, 139, 871–890. [Google Scholar] [CrossRef]
  6. Son, H.; Moon, A. Epithelial-mesenchymal Transition and Cell Invasion. Toxicol. Res. 2010, 26, 245–252. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Hofman, P.; Vouret-Craviari, V. Microbes-induced EMT at the crossroad of inflammation and cancer. Gut Microbes 2012, 3, 176–185. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Caven, L.T.; Brinkworth, A.J.; Carabeo, R.A. Pathogen-driven induction of a host transcriptome facilitating epithelial-to-mesenchymal transition. bioRxiv 2022, in press. [Google Scholar] [CrossRef]
  9. Champsi, J.; Young, L.S.; Bermudez, L.E. Production of TNF-alpha, IL-6 and TGF-beta, and expression of receptors for TNF-alpha and IL-6, during murine Mycobacterium avium infection. Immunology 1995, 84, 549–554. [Google Scholar] [PubMed]
  10. Reed, S.G. TGF-beta in infections and infectious diseases. Microbes Infect. 1999, 1, 1313–1325. [Google Scholar] [CrossRef]
  11. Silva, J.S.; Twardzik, D.R.; Reed, S.G. Regulation of Trypanosoma cruzi infections in vitro and in vivo by transforming growth factor beta (TGF-beta). J. Exp. Med. 1991, 174, 539–545. [Google Scholar] [CrossRef] [Green Version]
  12. Chapnick, D.A.; Warner, L.; Bernet, J.; Rao, T.; Liu, X. Partners in crime: The TGFβ and MAPK pathways in cancer progression. Cell Biosci. 2011, 1, 42. [Google Scholar] [CrossRef] [Green Version]
  13. Zhao, M.; Mishra, L.; Deng, C.-X. The role of TGF-β/SMAD4 signaling in cancer. Int. J. Biol. Sci. 2018, 14, 111–123. [Google Scholar] [CrossRef] [Green Version]
  14. Qin, J.; Li, R.; Raes, J.; Arumugam, M.; Burgdorf, K.S.; Manichanh, C.; Nielsen, T.; Pons, N.; Levenez, F.; Yamada, T.; et al. A human gut microbial gene catalogue established by metagenomic sequencing. Nature 2010, 464, 59–65. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Ley, R.E.; Peterson, D.A.; Gordon, J.I. Ecological and evolutionary forces shaping microbial diversity in the human intestine. Cell 2006, 124, 837–848. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Inagaki, H.; Suzuki, T.; Nomoto, K.; Yoshikai, Y. Increased susceptibility to primary infection with Listeria monocytogenes in germfree mice may be due to lack of accumulation of L-selectin+ CD44+ T cells in sites of inflammation. Infect. Immun. 1996, 64, 3280–3287. [Google Scholar] [CrossRef] [Green Version]
  17. Round, J.L.; Mazmanian, S.K. The gut microbiota shapes intestinal immune responses during health and disease. Nat. Rev. Immunol. 2009, 9, 313–323. [Google Scholar] [CrossRef]
  18. Wiertsema, S.P.; van Bergenhenegouwen, J.; Garssen, J.; Knippels, L.M.J. The Interplay between the Gut Microbiome and the Immune System in the Context of Infectious Diseases throughout Life and the Role of Nutrition in Optimizing Treatment Strategies. Nutrients 2021, 13, 886. [Google Scholar] [CrossRef]
  19. Human Microbiome Project Consortium. Structure, function and diversity of the healthy human microbiome. Nature 2012, 486, 207–214. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  20. Integrative HMP (iHMP) Research Network Consortium. The Integrative Human Microbiome Project: Dynamic analysis of microbiome-host omics profiles during periods of human health and disease. Cell Host Microbe 2014, 16, 276–289. [Google Scholar] [CrossRef] [Green Version]
  21. Morgillo, F.; Dallio, M.; Della Corte, C.M.; Gravina, A.G.; Viscardi, G.; Loguercio, C.; Ciardiello, F.; Federico, A. Carcinogenesis as a Result of Multiple Inflammatory and Oxidative Hits: A Comprehensive Review from Tumor Microenvironment to Gut Microbiota. Neoplasia 2018, 20, 721–733. [Google Scholar] [CrossRef] [PubMed]
  22. Bhatt, A.P.; Redinbo, M.R.; Bultman, S.J. The role of the microbiome in cancer development and therapy. CA Cancer J. Clin. 2017, 67, 326–344. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Coker, O.O.; Wu, W.K.K.; Wong, S.H.; Sung, J.J.Y.; Yu, J. Altered Gut Archaea Composition and Interaction with Bacteria Are Associated With Colorectal Cancer. Gastroenterology 2020, 159, 1459–1470.e1455. [Google Scholar] [CrossRef] [PubMed]
  24. Feng, Q.; Liang, S.; Jia, H.; Stadlmayr, A.; Tang, L.; Lan, Z.; Zhang, D.; Xia, H.; Xu, X.; Jie, Z.; et al. Gut microbiome development along the colorectal adenoma–carcinoma sequence. Nat. Commun. 2015, 6, 6528. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Liu, W.; Zhang, X.; Xu, H.; Li, S.; Lau, H.C.-H.; Chen, Q.; Zhang, B.; Zhao, L.; Chen, H.; Sung, J.J.-Y.; et al. Microbial Community Heterogeneity Within Colorectal Neoplasia and its Correlation With Colorectal Carcinogenesis. Gastroenterology 2021, 160, 2395–2408. [Google Scholar] [CrossRef]
  26. Nakatsu, G.; Zhou, H.; Wu, W.K.K.; Wong, S.H.; Coker, O.O.; Dai, Z.; Li, X.; Szeto, C.-H.; Sugimura, N.; Lam, T.Y.-T.; et al. Alterations in Enteric Virome Are Associated with Colorectal Cancer and Survival Outcomes. Gastroenterology 2018, 155, 529–541.e525. [Google Scholar] [CrossRef] [Green Version]
  27. Thomas, A.M.; Manghi, P.; Asnicar, F.; Pasolli, E.; Armanini, F.; Zolfo, M.; Beghini, F.; Manara, S.; Karcher, N.; Pozzi, C.; et al. Metagenomic analysis of colorectal cancer datasets identifies cross-cohort microbial diagnostic signatures and a link with choline degradation. Nat. Med. 2019, 25, 667–678. [Google Scholar] [CrossRef] [Green Version]
  28. Wang, Y.; Wan, X.; Wu, X.; Zhang, C.; Liu, J.; Hou, S. Eubacterium rectale contributes to colorectal cancer initiation via promoting colitis. Gut Pathog. 2021, 13, 2. [Google Scholar] [CrossRef]
  29. Wirbel, J.; Pyl, P.T.; Kartal, E.; Zych, K.; Kashani, A.; Milanese, A.; Fleck, J.S.; Voigt, A.Y.; Palleja, A.; Ponnudurai, R.; et al. Meta-analysis of fecal metagenomes reveals global microbial signatures that are specific for colorectal cancer. Nat. Med. 2019, 25, 679–689. [Google Scholar] [CrossRef] [Green Version]
  30. Yang, T.-W.; Lee, W.-H.; Tu, S.-J.; Huang, W.-C.; Chen, H.-M.; Sun, T.-H.; Tsai, M.-C.; Wang, C.-C.; Chen, H.-Y.; Huang, C.-C.; et al. Enterotype-based Analysis of Gut Microbiota along the Conventional Adenoma-Carcinoma Colorectal Cancer Pathway. Sci. Rep. 2019, 9, 10923. [Google Scholar] [CrossRef] [Green Version]
  31. Yang, Y.; Misra, B.B.; Liang, L.; Bi, D.; Weng, W.; Wu, W.; Cai, S.; Qin, H.; Goel, A.; Li, X.; et al. Integrated microbiome and metabolome analysis reveals a novel interplay between commensal bacteria and metabolites in colorectal cancer. Theranostics 2019, 9, 4101–4114. [Google Scholar] [CrossRef] [PubMed]
  32. Zackular, J.P.; Rogers, M.A.M.; Ruffin, M.T.t.; Schloss, P.D. The human gut microbiome as a screening tool for colorectal cancer. Cancer Prev. Res. 2014, 7, 1112–1121. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Fernández, M.F.; Reina-Pérez, I.; Astorga, J.M.; Rodríguez-Carrillo, A.; Plaza-Díaz, J.; Fontana, L. Breast Cancer and Its Relationship with the Microbiota. Int. J. Environ. Res. Public Health 2018, 15, 1747. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Gui, Q.F.; Lu, H.F.; Zhang, C.X.; Xu, Z.R.; Yang, Y.H. Well-balanced commensal microbiota contributes to anti-cancer response in a lung cancer mouse model. Genet. Mol. Res. 2015, 14, 5642–5651. [Google Scholar] [CrossRef] [PubMed]
  35. Raza, M.H.; Gul, K.; Arshad, A.; Riaz, N.; Waheed, U.; Rauf, A.; Aldakheel, F.; Alduraywish, S.; Rehman, M.U.; Abdullah, M.; et al. Microbiota in cancer development and treatment. J. Cancer Res. Clin. Oncol. 2019, 145, 49–63. [Google Scholar] [CrossRef]
  36. Routy, B.; Le Chatelier, E.; Derosa, L.; Duong, C.P.M.; Alou, M.T.; Daillère, R.; Fluckiger, A.; Messaoudene, M.; Rauber, C.; Roberti, M.P.; et al. Gut microbiome influences efficacy of PD-1-based immunotherapy against epithelial tumors. Science 2018, 359, 91–97. [Google Scholar] [CrossRef] [Green Version]
  37. Zhang, W.Q.; Zhao, S.K.; Luo, J.W.; Dong, X.P.; Hao, Y.T.; Li, H.; Shan, L.; Zhou, Y.; Shi, H.B.; Zhang, Z.Y.; et al. Alterations of fecal bacterial communities in patients with lung cancer. Am. J. Transl. Res. 2018, 10, 3171–3185. [Google Scholar]
  38. Banerjee, S.; Tian, T.; Wei, Z.; Shih, N.; Feldman, M.D.; Alwine, J.C.; Coukos, G.; Robertson, E.S. The ovarian cancer oncobiome. Oncotarget 2017, 8, 36225–36245. [Google Scholar] [CrossRef] [Green Version]
  39. Łaniewski, P.; Barnes, D.; Goulder, A.; Cui, H.; Roe, D.J.; Chase, D.M.; Herbst-Kralovetz, M.M. Linking cervicovaginal immune signatures, HPV and microbiota composition in cervical carcinogenesis in non-Hispanic and Hispanic women. Sci. Rep. 2018, 8, 7593. [Google Scholar] [CrossRef]
  40. Łaniewski, P.; Ilhan, Z.E.; Herbst-Kralovetz, M.M. The microbiome and gynaecological cancer development, prevention and therapy. Nat. Rev. Urol. 2020, 17, 232–250. [Google Scholar] [CrossRef]
  41. Zhou, B.; Sun, C.; Huang, J.; Xia, M.; Guo, E.; Li, N.; Lu, H.; Shan, W.; Wu, Y.; Li, Y.; et al. The biodiversity Composition of Microbiome in Ovarian Carcinoma Patients. Sci. Rep. 2019, 9, 1691. [Google Scholar] [CrossRef] [PubMed]
  42. Rubinstein, M.R.; Wang, X.; Liu, W.; Hao, Y.; Cai, G.; Han, Y.W. Fusobacterium nucleatum promotes colorectal carcinogenesis by modulating E-cadherin/β-catenin signaling via its FadA adhesin. Cell Host Microbe 2013, 14, 195–206. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Chen, L.; Gibbons, D.L.; Goswami, S.; Cortez, M.A.; Ahn, Y.H.; Byers, L.A.; Zhang, X.; Yi, X.; Dwyer, D.; Lin, W.; et al. Metastasis is regulated via microRNA-200/ZEB1 axis control of tumour cell PD-L1 expression and intratumoral immunosuppression. Nat. Commun. 2014, 5, 5241. [Google Scholar] [CrossRef] [PubMed]
  44. Sacdalan, D.B.; Lucero, J.A. The Association Between Inflammation and Immunosuppression: Implications for ICI Biomarker Development. OncoTargets Ther. 2021, 14, 2053–2064. [Google Scholar] [CrossRef]
  45. Kostic, A.D.; Chun, E.; Robertson, L.; Glickman, J.N.; Gallini, C.A.; Michaud, M.; Clancy, T.E.; Chung, D.C.; Lochhead, P.; Hold, G.L.; et al. Fusobacterium nucleatum potentiates intestinal tumorigenesis and modulates the tumor-immune microenvironment. Cell Host Microbe 2013, 14, 207–215. [Google Scholar] [CrossRef] [Green Version]
  46. Hooi, J.K.Y.; Lai, W.Y.; Ng, W.K.; Suen, M.M.Y.; Underwood, F.E.; Tanyingoh, D.; Malfertheiner, P.; Graham, D.Y.; Wong, V.W.S.; Wu, J.C.Y.; et al. Global Prevalence of Helicobacter pylori Infection: Systematic Review and Meta-Analysis. Gastroenterology 2017, 153, 420–429. [Google Scholar] [CrossRef] [Green Version]
  47. Mitchell, H.M. The epidemiology of Helicobacter pylori. Curr. Top. Microbiol. Immunol. 1999, 241, 11–30. [Google Scholar] [CrossRef]
  48. Smith, M.F., Jr.; Mitchell, A.; Li, G.; Ding, S.; Fitzmaurice, A.M.; Ryan, K.; Crowe, S.; Goldberg, J.B. Toll-like receptor (TLR) 2 and TLR5, but not TLR4, are required for Helicobacter pylori-induced NF-kappa B activation and chemokine expression by epithelial cells. J. Biol. Chem. 2003, 278, 32552–32560. [Google Scholar] [CrossRef] [Green Version]
  49. Papini, E.; Satin, B.; Norais, N.; de Bernard, M.; Telford, J.L.; Rappuoli, R.; Montecucco, C. Selective increase of the permeability of polarized epithelial cell monolayers by Helicobacter pylori vacuolating toxin. J. Clin. Investig. 1998, 102, 813–820. [Google Scholar] [CrossRef] [Green Version]
  50. Amieva, M.R.; Vogelmann, R.; Covacci, A.; Tompkins, L.S.; Nelson, W.J.; Falkow, S. Disruption of the epithelial apical-junctional complex by Helicobacter pylori CagA. Science 2003, 300, 1430–1434. [Google Scholar] [CrossRef] [Green Version]
  51. Murata-Kamiya, N.; Kurashima, Y.; Teishikata, Y.; Yamahashi, Y.; Saito, Y.; Higashi, H.; Aburatani, H.; Akiyama, T.; Peek, R.M., Jr.; Azuma, T.; et al. Helicobacter pylori CagA interacts with E-cadherin and deregulates the beta-catenin signal that promotes intestinal transdifferentiation in gastric epithelial cells. Oncogene 2007, 26, 4617–4626. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Brandt, S.; Kwok, T.; Hartig, R.; König, W.; Backert, S. NF-kappaB activation and potentiation of proinflammatory responses by the Helicobacter pylori CagA protein. Proc. Natl. Acad. Sci. USA 2005, 102, 9300–9305. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Yin, Y.; Grabowska, A.M.; Clarke, P.A.; Whelband, E.; Robinson, K.; Argent, R.H.; Tobias, A.; Kumari, R.; Atherton, J.C.; Watson, S.A. Helicobacter pylori potentiates epithelial: Mesenchymal transition in gastric cancer: Links to soluble HB-EGF, gastrin and matrix metalloproteinase-7. Gut 2010, 59, 1037–1045. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Murphy, R.; Palm, M.; Mustonen, V.; Warringer, J.; Farewell, A.; Parts, L.; Moradigaravand, D. Genomic Epidemiology and Evolution of Escherichia coli in Wild Animals in Mexico. mSphere 2021, 6, e00738-20. [Google Scholar] [CrossRef] [PubMed]
  55. König, J.; Wells, J.; Cani, P.D.; García-Ródenas, C.L.; MacDonald, T.; Mercenier, A.; Whyte, J.; Troost, F.; Brummer, R.-J. Human Intestinal Barrier Function in Health and Disease. Clin. Transl. Gastroenterol. 2016, 7, e196. [Google Scholar] [CrossRef]
  56. Bétis, F.; Brest, P.; Hofman, V.; Guignot, J.; Bernet-Camard, M.F.; Rossi, B.; Servin, A.; Hofman, P. The Afa/Dr adhesins of diffusely adhering Escherichia coli stimulate interleukin-8 secretion, activate mitogen-activated protein kinases, and promote polymorphonuclear transepithelial migration in T84 polarized epithelial cells. Infect. Immun. 2003, 71, 1068–1074. [Google Scholar] [CrossRef] [Green Version]
  57. Bétis, F.; Brest, P.; Hofman, V.; Guignot, J.; Kansau, I.; Rossi, B.; Servin, A.; Hofman, P. Afa/Dr diffusely adhering Escherichia coli infection in T84 cell monolayers induces increased neutrophil transepithelial migration, which in turn promotes cytokine-dependent upregulation of decay-accelerating factor (CD55), the receptor for Afa/Dr adhesins. Infect. Immun. 2003, 71, 1774–1783. [Google Scholar] [CrossRef] [Green Version]
  58. Cane, G.; Moal, V.L.; Pagès, G.; Servin, A.L.; Hofman, P.; Vouret-Craviari, V. Up-regulation of intestinal vascular endothelial growth factor by Afa/Dr diffusely adhering Escherichia coli. PLoS ONE 2007, 2, e1359. [Google Scholar] [CrossRef] [Green Version]
  59. Cane, G.; Ginouvès, A.; Marchetti, S.; Buscà, R.; Pouysségur, J.; Berra, E.; Hofman, P.; Vouret-Craviari, V. HIF-1alpha mediates the induction of IL-8 and VEGF expression on infection with Afa/Dr diffusely adhering E. coli and promotes EMT-like behaviour. Cell. Microbiol. 2010, 12, 640–653. [Google Scholar] [CrossRef]
  60. Park, B.S.; Lee, J.-O. Recognition of lipopolysaccharide pattern by TLR4 complexes. Exp. Mol. Med. 2013, 45, e66. [Google Scholar] [CrossRef] [Green Version]
  61. Zhao, L.; Yang, R.; Cheng, L.; Wang, M.; Jiang, Y.; Wang, S. LPS-induced epithelial-mesenchymal transition of intrahepatic biliary epithelial cells. J. Surg. Res. 2011, 171, 819–825. [Google Scholar] [CrossRef] [PubMed]
  62. Miyazono, K. Transforming growth factor-beta signaling in epithelial-mesenchymal transition and progression of cancer. Proc. Jpn. Acad. Ser. B Phys. Biol. Sci. 2009, 85, 314–323. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Honko, A.N.; Mizel, S.B. Effects of flagellin on innate and adaptive immunity. Immunol. Res. 2005, 33, 83–101. [Google Scholar] [CrossRef]
  64. Franchi, L.; Park, J.H.; Shaw, M.H.; Marina-Garcia, N.; Chen, G.; Kim, Y.G.; Núñez, G. Intracellular NOD-like receptors in innate immunity, infection and disease. Cell. Microbiol. 2008, 10, 1–8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Yang, J.J.; Wang, D.D.; Sun, T.Y. Flagellin of Pseudomonas aeruginosa induces transforming growth factor beta 1 expression in normal bronchial epithelial cells through mitogen activated protein kinase cascades. Chin. Med. J. 2011, 124, 599–605. [Google Scholar] [PubMed]
  66. Ferrand, A.; Al Nabhani, Z.; Tapias, N.S.; Mas, E.; Hugot, J.-P.; Barreau, F. NOD2 Expression in Intestinal Epithelial Cells Protects Toward the Development of Inflammation and Associated Carcinogenesis. Cell. Mol. Gastroenterol. Hepatol. 2019, 7, 357–369. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Dominguez, C.; David, J.M.; Palena, C. Epithelial-mesenchymal transition and inflammation at the site of the primary tumor. Semin. Cancer Biol. 2017, 47, 177–184. [Google Scholar] [CrossRef] [PubMed]
  68. Sun, L.; Suo, C.; Li, S.T.; Zhang, H.; Gao, P. Metabolic reprogramming for cancer cells and their microenvironment: Beyond the Warburg Effect. Biochim. Biophys. Acta Rev. Cancer 2018, 1870, 51–66. [Google Scholar] [CrossRef] [PubMed]
  69. Kovács, T.; Mikó, E.; Vida, A.; Sebő, É.; Toth, J.; Csonka, T.; Boratkó, A.; Ujlaki, G.; Lente, G.; Kovács, P.; et al. Cadaverine, a metabolite of the microbiome, reduces breast cancer aggressiveness through trace amino acid receptors. Sci. Rep. 2019, 9, 1300. [Google Scholar] [CrossRef] [Green Version]
  70. Mikó, E.; Vida, A.; Kovács, T.; Ujlaki, G.; Trencsényi, G.; Márton, J.; Sári, Z.; Kovács, P.; Boratkó, A.; Hujber, Z.; et al. Lithocholic acid, a bacterial metabolite reduces breast cancer cell proliferation and aggressiveness. Biochim. Biophys. Acta Bioenerg. 2018, 1859, 958–974. [Google Scholar] [CrossRef]
  71. François, A.; Grebert, D.; Rhimi, M.; Mariadassou, M.; Naudon, L.; Rabot, S.; Meunier, N. Olfactory epithelium changes in germfree mice. Sci. Rep. 2016, 6, 24687. [Google Scholar] [CrossRef] [PubMed]
  72. Rawls, M.; Ellis, A.K. The microbiome of the nose. Ann. Allergy Asthma Immunol. 2019, 122, 17–24. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Salzano, F.A.; Marino, L.; Salzano, G.; Botta, R.M.; Cascone, G.; D’Agostino Fiorenza, U.; Selleri, C.; Casolaro, V. Microbiota Composition and the Integration of Exogenous and Endogenous Signals in Reactive Nasal Inflammation. J. Immunol. Res. 2018, 2018, 2724951. [Google Scholar] [CrossRef]
  74. Evans, S.E.; Xu, Y.; Tuvim, M.J.; Dickey, B.F. Inducible innate resistance of lung epithelium to infection. Annu. Rev. Physiol. 2010, 72, 413–435. [Google Scholar] [CrossRef] [Green Version]
  75. Ziesemer, S.; Eiffler, I.; Schönberg, A.; Müller, C.; Hochgräfe, F.; Beule, A.G.; Hildebrandt, J.P. Staphylococcus aureus α-Toxin Induces Actin Filament Remodeling in Human Airway Epithelial Model Cells. Am. J. Respir. Cell Mol. Biol. 2018, 58, 482–491. [Google Scholar] [CrossRef]
  76. Patou, J.; Gevaert, P.; Van Zele, T.; Holtappels, G.; van Cauwenberge, P.; Bachert, C. Staphylococcus aureus enterotoxin B, protein A, and lipoteichoic acid stimulations in nasal polyps. J. Allergy Clin. Immunol. 2008, 121, 110–115. [Google Scholar] [CrossRef]
  77. Meng, J.; Zhou, P.; Liu, Y.; Liu, F.; Yi, X.; Liu, S.; Holtappels, G.; Bachert, C.; Zhang, N. The development of nasal polyp disease involves early nasal mucosal inflammation and remodelling. PLoS ONE 2013, 8, e82373. [Google Scholar] [CrossRef] [Green Version]
  78. Dickson, R.P.; Erb-Downward, J.R.; Martinez, F.J.; Huffnagle, G.B. The Microbiome and the Respiratory Tract. Annu. Rev. Physiol. 2016, 78, 481–504. [Google Scholar] [CrossRef] [Green Version]
  79. Mur, L.A.; Huws, S.A.; Cameron, S.J.; Lewis, P.D.; Lewis, K.E. Lung cancer: A new frontier for microbiome research and clinical translation. Ecancermedicalscience 2018, 12, 866. [Google Scholar] [CrossRef]
  80. Mao, Q.; Jiang, F.; Yin, R.; Wang, J.; Xia, W.; Dong, G.; Ma, W.; Yang, Y.; Xu, L.; Hu, J. Interplay between the lung microbiome and lung cancer. Cancer Lett. 2018, 415, 40–48. [Google Scholar] [CrossRef] [PubMed]
  81. Jin, C.; Lagoudas, G.K.; Zhao, C.; Bullman, S.; Bhutkar, A.; Hu, B.; Ameh, S.; Sandel, D.; Liang, X.S.; Mazzilli, S.; et al. Commensal Microbiota Promote Lung Cancer Development via γδ T Cells. Cell 2019, 176, 998–1013.e16. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Zhuang, H.; Cheng, L.; Wang, Y.; Zhang, Y.K.; Zhao, M.F.; Liang, G.D.; Zhang, M.C.; Li, Y.G.; Zhao, J.B.; Gao, Y.N.; et al. Dysbiosis of the Gut Microbiome in Lung Cancer. Front. Cell. Infect. Microbiol. 2019, 9, 112. [Google Scholar] [CrossRef]
  83. Botticelli, A.; Putignani, L.; Zizzari, I.; Chierico, F.D.; Reddel, S.; Pietro, F.D.; Quagliarello, A.; Onesti, C.E.; Raffaele, G.; Mazzuca, F.; et al. Changes of microbiome profile during nivolumab treatment in NSCLC patients. J. Clin. Oncol. 2018, 36, e15020. [Google Scholar] [CrossRef]
  84. Gui, Q.; Li, H.; Wang, A.; Zhao, X.; Tan, Z.; Chen, L.; Xu, K.; Xiao, C. The association between gut butyrate-producing bacteria and non-small-cell lung cancer. J. Clin. Lab. Anal. 2020, 34, e23318. [Google Scholar] [CrossRef] [Green Version]
  85. Liu, F.; Li, J.; Guan, Y.; Lou, Y.; Chen, H.; Xu, M.; Deng, D.; Chen, J.; Ni, B.; Zhao, L.; et al. Dysbiosis of the Gut Microbiome is associated with Tumor Biomarkers in Lung Cancer. Int. J. Biol. Sci. 2019, 15, 2381–2392. [Google Scholar] [CrossRef]
  86. Zheng, Y.; Fang, Z.; Xue, Y.; Zhang, J.; Zhu, J.; Gao, R.; Yao, S.; Ye, Y.; Wang, S.; Lin, C.; et al. Specific gut microbiome signature predicts the early-stage lung cancer. Gut Microbes 2020, 11, 1030–1042. [Google Scholar] [CrossRef]
  87. Yang, J.J.; Yu, D.; Xiang, Y.B.; Blot, W.; White, E.; Robien, K.; Sinha, R.; Park, Y.; Takata, Y.; Lazovich, D.; et al. Association of Dietary Fiber and Yogurt Consumption With Lung Cancer Risk: A Pooled Analysis. JAMA Oncol. 2020, 6, e194107. [Google Scholar] [CrossRef]
  88. Bai, Y.; Shen, W.; Zhu, M.; Zhang, L.; Wei, Y.; Tang, H.; Zhao, J. Combined detection of estrogen and tumor markers is an important reference factor in the diagnosis and prognosis of lung cancer. J. Cell. Biochem. 2019, 120, 105–114. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  89. Gill, N.; Wlodarska, M.; Finlay, B.B. The future of mucosal immunology: Studying an integrated system-wide organ. Nat. Immunol. 2010, 11, 558–560. [Google Scholar] [CrossRef] [PubMed]
  90. Renz, H.; Brandtzaeg, P.; Hornef, M. The impact of perinatal immune development on mucosal homeostasis and chronic inflammation. Nat. Rev. Immunol. 2011, 12, 9–23. [Google Scholar] [CrossRef]
  91. Bingula, R.; Filaire, M.; Radosevic-Robin, N.; Bey, M.; Berthon, J.Y.; Bernalier-Donadille, A.; Vasson, M.P.; Filaire, E. Desired Turbulence? Gut-Lung Axis, Immunity, and Lung Cancer. J. Oncol. 2017, 2017, 5035371. [Google Scholar] [CrossRef] [PubMed]
  92. Hooper, L.V.; Littman, D.R.; Macpherson, A.J. Interactions between the microbiota and the immune system. Science 2012, 336, 1268–1273. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Mazmanian, S.K.; Liu, C.H.; Tzianabos, A.O.; Kasper, D.L. An immunomodulatory molecule of symbiotic bacteria directs maturation of the host immune system. Cell 2005, 122, 107–118. [Google Scholar] [CrossRef] [Green Version]
  94. Samuelson, D.R.; Welsh, D.A.; Shellito, J.E. Regulation of lung immunity and host defense by the intestinal microbiota. Front. Microbiol. 2015, 6, 1085. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Dumas, A.; Bernard, L.; Poquet, Y.; Lugo-Villarino, G.; Neyrolles, O. The role of the lung microbiota and the gut-lung axis in respiratory infectious diseases. Cell. Microbiol. 2018, 20, e12966. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Schuijt, T.J.; Lankelma, J.M.; Scicluna, B.P.; de Sousa e Melo, F.; Roelofs, J.J.; de Boer, J.D.; Hoogendijk, A.J.; de Beer, R.; de Vos, A.; Belzer, C.; et al. The gut microbiota plays a protective role in the host defence against pneumococcal pneumonia. Gut 2016, 65, 575–583. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Kim, M.; Gu, B.; Madison, M.C.; Song, H.W.; Norwood, K.; Hill, A.A.; Wu, W.J.; Corry, D.; Kheradmand, F.; Diehl, G.E. Cigarette Smoke Induces Intestinal Inflammation via a Th17 Cell-Neutrophil Axis. Front. Immunol. 2019, 10, 75. [Google Scholar] [CrossRef] [Green Version]
  98. Boursi, B.; Mamtani, R.; Haynes, K.; Yang, Y.X. Recurrent antibiotic exposure may promote cancer formation--Another step in understanding the role of the human microbiota? Eur. J. Cancer 2015, 51, 2655–2664. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  99. Druzhinin, V.G.; Matskova, L.V.; Fucic, A. Induction and modulation of genotoxicity by the bacteriome in mammals. Mutat. Res./Rev. Mutat. Res. 2018, 776, 70–77. [Google Scholar] [CrossRef]
  100. Louis, P.; Hold, G.L.; Flint, H.J. The gut microbiota, bacterial metabolites and colorectal cancer. Nat. Rev. Microbiol. 2014, 12, 661–672. [Google Scholar] [CrossRef] [PubMed]
  101. Van Raay, T.; Allen-Vercoe, E. Microbial Interactions and Interventions in Colorectal Cancer. Microbiol. Spectr. 2017, 5, 99–130. [Google Scholar] [CrossRef] [PubMed]
  102. Zhao, H.; Chu, M.; Huang, Z.; Yang, X.; Ran, S.; Hu, B.; Zhang, C.; Liang, J. Variations in oral microbiota associated with oral cancer. Sci. Rep. 2017, 7, 11773. [Google Scholar] [CrossRef] [PubMed]
  103. Lee, W.H.; Chen, H.M.; Yang, S.F.; Liang, C.; Peng, C.Y.; Lin, F.M.; Tsai, L.L.; Wu, B.C.; Hsin, C.H.; Chuang, C.Y.; et al. Bacterial alterations in salivary microbiota and their association in oral cancer. Sci. Rep. 2017, 7, 16540. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Whitmore, S.E.; Lamont, R.J. Oral bacteria and cancer. PLoS Pathog. 2014, 10, e1003933. [Google Scholar] [CrossRef]
  105. Nagy, K.N.; Sonkodi, I.; Szöke, I.; Nagy, E.; Newman, H.N. The microflora associated with human oral carcinomas. Oral Oncol. 1998, 34, 304–308. [Google Scholar] [CrossRef]
  106. Karpiński, T.M. Role of oral microbiota in cancer development. Microorganisms 2019, 7, 20. [Google Scholar] [CrossRef] [Green Version]
  107. Ha, N.H.; Woo, B.H.; Kim, D.J.; Ha, E.S.; Choi, J.I.; Kim, S.J.; Park, B.S.; Lee, J.H.; Park, H.R. Prolonged and repetitive exposure to Porphyromonas gingivalis increases aggressiveness of oral cancer cells by promoting acquisition of cancer stem cell properties. Tumour Biol. 2015, 36, 9947–9960. [Google Scholar] [CrossRef]
  108. Sztukowska, M.N.; Ojo, A.; Ahmed, S.; Carenbauer, A.L.; Wang, Q.; Shumway, B.; Jenkinson, H.F.; Wang, H.; Darling, D.S.; Lamont, R.J. Porphyromonas gingivalis initiates a mesenchymal-like transition through ZEB1 in gingival epithelial cells. Cell. Microbiol. 2016, 18, 844–858. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  109. Snider, E.J.; Compres, G.; Freedberg, D.E.; Khiabanian, H.; Nobel, Y.R.; Stump, S.; Uhlemann, A.-C.; Lightdale, C.J.; Abrams, J.A. Alterations to the esophageal microbiome associated with progression from Barrett’s esophagus to esophageal adenocarcinoma. Cancer Epidemiol. Prev. Biomark. 2019, 28, 1687–1693. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  110. Wang, Q.; Rao, Y.; Guo, X.; Liu, N.; Liu, S.; Wen, P.; Li, S.; Li, Y. Oral microbiome in patients with oesophageal squamous cell carcinoma. Sci. Rep. 2019, 9, 19055. [Google Scholar] [CrossRef] [Green Version]
  111. Chen, X.; Winckler, B.; Lu, M.; Cheng, H.; Yuan, Z.; Yang, Y.; Jin, L.; Ye, W. Oral microbiota and risk for esophageal squamous cell carcinoma in a high-risk area of China. PLoS ONE 2015, 10, e0143603. [Google Scholar] [CrossRef] [PubMed]
  112. Blackett, K.; Siddhi, S.; Cleary, S.; Steed, H.; Miller, M.; Macfarlane, S.; Macfarlane, G.; Dillon, J. Oesophageal bacterial biofilm changes in gastro-oesophageal reflux disease, Barrett’s and oesophageal carcinoma: Association or causality? Aliment. Pharmacol. Ther. 2013, 37, 1084–1092. [Google Scholar] [CrossRef] [PubMed]
  113. Cass, S.; Hamilton, C.; Miller, A.; Jupiter, D.; Khanipov, K.; Booth, A.; Pyles, R.; Krill, T.; Reep, G.; Okereke, I. Novel ex vivo model to examine the mechanism and relationship of esophageal microbiota and disease. Biomedicines 2021, 9, 142. [Google Scholar] [CrossRef] [PubMed]
  114. Elliott, D.R.F.; Walker, A.W.; O’Donovan, M.; Parkhill, J.; Fitzgerald, R.C. A non-endoscopic device to sample the oesophageal microbiota: A case-control study. Lancet Gastroenterol. Hepatol. 2017, 2, 32–42. [Google Scholar] [CrossRef] [Green Version]
  115. Li, D.; He, R.; Hou, G.; Ming, W.; Fan, T.; Chen, L.; Zhang, L.; Jiang, W.; Wang, W.; Lu, Z. Characterization of the esophageal microbiota and prediction of the metabolic pathways involved in esophageal cancer. Front. Cell. Infect. Microbiol. 2020, 10, 268. [Google Scholar] [CrossRef]
  116. Shao, D.; Vogtmann, E.; Liu, A.; Qin, J.; Chen, W.; Abnet, C.C.; Wei, W. Microbial characterization of esophageal squamous cell carcinoma and gastric cardia adenocarcinoma from a high-risk region of China. Cancer 2019, 125, 3993–4002. [Google Scholar] [CrossRef]
  117. Peters, B.A.; Wu, J.; Pei, Z.; Yang, L.; Purdue, M.P.; Freedman, N.D.; Jacobs, E.J.; Gapstur, S.M.; Hayes, R.B.; Ahn, J. Oral microbiome composition reflects prospective risk for esophageal cancers. Cancer Res. 2017, 77, 6777–6787. [Google Scholar] [CrossRef] [Green Version]
  118. Narikiyo, M.; Tanabe, C.; Yamada, Y.; Igaki, H.; Tachimori, Y.; Kato, H.; Muto, M.; Montesano, R.; Sakamoto, H.; Nakajima, Y. Frequent and preferential infection of Treponema denticola, Streptococcus mitis, and Streptococcus anginosus in esophageal cancers. Cancer Sci. 2004, 95, 569–574. [Google Scholar] [CrossRef]
  119. Lopetuso, L.R.; Severgnini, M.; Pecere, S.; Ponziani, F.R.; Boskoski, I.; Larghi, A.; Quaranta, G.; Masucci, L.; Ianiro, G.; Camboni, T. Esophageal microbiome signature in patients with Barrett’s esophagus and esophageal adenocarcinoma. PLoS ONE 2020, 15, e0231789. [Google Scholar] [CrossRef]
  120. Liu, Y.; Lin, Z.; Lin, Y.; Chen, Y.; Peng, X.-E.; He, F.; Liu, S.; Yan, S.; Huang, L.; Lu, W. Streptococcus and Prevotella are associated with the prognosis of oesophageal squamous cell carcinoma. J. Med. Microbiol. 2018, 67, 1058–1068. [Google Scholar] [CrossRef]
  121. Chen, M.-F.; Lu, M.-S.; Hsieh, C.-C.; Chen, W.-C. Porphyromonas gingivalis promotes tumor progression in esophageal squamous cell carcinoma. Cell. Oncol. 2021, 44, 373–384. [Google Scholar] [CrossRef] [PubMed]
  122. Kaakoush, N.O.; Lecomte, V.; Maloney, C.A.; Morris, M.J. Cross-talk among metabolic parameters, esophageal microbiota, and host gene expression following chronic exposure to an obesogenic diet. Sci. Rep. 2017, 7, 45753. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Münch, N.S.; Fang, H.-Y.; Ingermann, J.; Maurer, H.C.; Anand, A.; Kellner, V.; Sahm, V.; Wiethaler, M.; Baumeister, T.; Wein, F. High-fat diet accelerates carcinogenesis in a mouse model of Barrett’s esophagus via interleukin 8 and alterations to the gut microbiome. Gastroenterology 2019, 157, 492–506.e2. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Pham, V.T.; Fehlbaum, S.; Seifert, N.; Richard, N.; Bruins, M.J.; Sybesma, W.; Rehman, A.; Steinert, R.E. Effects of colon-targeted vitamins on the composition and metabolic activity of the human gut microbiome—A pilot study. Gut Microbes 2021, 13, 1875774. [Google Scholar] [CrossRef]
  125. Pan, F.; Xu, X.; Zhang, L.-L.; Luo, H.-J.; Chen, Y.; Long, L.; Wang, X.; Zhuang, P.-T.; Li, E.-M.; Xu, L.-Y. Correction: Dietary riboflavin deficiency induces genomic instability of esophageal squamous cells that is associated with gut microbiota dysbiosis in rats. Food Funct. 2020, 11, 10979. [Google Scholar] [CrossRef]
  126. Pan, F.; Zhang, L.-L.; Luo, H.-J.; Chen, Y.; Long, L.; Wang, X.; Zhuang, P.-T.; Li, E.-M.; Xu, L.-Y. Dietary riboflavin deficiency induces ariboflavinosis and esophageal epithelial atrophy in association with modification of gut microbiota in rats. Eur. J. Nutr. 2021, 60, 807–820. [Google Scholar] [CrossRef]
  127. Chow, W.-H.; Blaser, M.J.; Blot, W.J.; Gammon, M.D.; Vaughan, T.L.; Risch, H.A.; Perez-Perez, G.I.; Schoenberg, J.B.; Stanford, J.L.; Rotterdam, H. An inverse relation between cagA+ strains of Helicobacter pylori infection and risk of esophageal and gastric cardia adenocarcinoma. Cancer Res. 1998, 58, 588–590. [Google Scholar]
  128. Nie, S.; Chen, T.; Yang, X.; Huai, P.; Lu, M. Association of h elicobacter pylori infection with esophageal adenocarcinoma and squamous cell carcinoma: A meta-analysis. Dis. Esophagus 2014, 27, 645–653. [Google Scholar] [CrossRef]
  129. Peek, R.M.; Blaser, M.J. Helicobacter pylori and gastrointestinal tract adenocarcinomas. Nat. Rev. Cancer 2002, 2, 28–37. [Google Scholar] [CrossRef]
  130. Kumar, S.; Metz, D.C.; Ginsberg, G.G.; Kaplan, D.E.; Goldberg, D.S. Oesophageal and proximal gastric adenocarcinomas are rare after detection of Helicobacter pylori infection. Aliment. Pharmacol. Ther. 2020, 51, 781–788. [Google Scholar] [CrossRef]
  131. Doorakkers, E.; Lagergren, J.; Santoni, G.; Engstrand, L.; Brusselaers, N. Helicobacter pylori eradication treatment and the risk of Barrett’s esophagus and esophageal adenocarcinoma. Helicobacter 2020, 25, e12688. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  132. Ferreira, R.M.; Pereira-Marques, J.; Pinto-Ribeiro, I.; Costa, J.L.; Carneiro, F.; Machado, J.C.; Figueiredo, C. Gastric microbial community profiling reveals a dysbiotic cancer-associated microbiota. Gut 2018, 67, 226–236. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Amieva, M.; Peek Jr, R.M. Pathobiology of Helicobacter pylori–induced gastric cancer. Gastroenterology 2016, 150, 64–78. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Hansen, A.; Johannesen, T.B.; Spiegelhauer, M.; Kupcinskas, J.; Urba, M.; Skieceviciene, J.; Jonaitis, L.; Frandsen, T.; Kupcinskas, L.; Fuursted, K. Distinct composition and distribution of the gastric mycobiota observed between dyspeptic and gastric cancer patients evaluated from gastric biopsies. Microb. Health Dis. 2020, 2, e340. [Google Scholar]
  135. Spiegelhauer, M.R.; Kupcinskas, J.; Johannesen, T.B.; Urba, M.; Skieceviciene, J.; Jonaitis, L.; Frandsen, T.H.; Kupcinskas, L.; Fuursted, K.; Andersen, L.P. Transient and persistent gastric microbiome: Adherence of bacteria in gastric cancer and dyspeptic patient biopsies after washing. J. Clin. Med. 2020, 9, 1882. [Google Scholar] [CrossRef] [PubMed]
  136. Yu, G.; Torres, J.; Hu, N.; Medrano-Guzman, R.; Herrera-Goepfert, R.; Humphrys, M.S.; Wang, L.; Wang, C.; Ding, T.; Ravel, J. Molecular characterization of the human stomach microbiota in gastric cancer patients. Front. Cell. Infect. Microbiol. 2017, 7, 302. [Google Scholar] [CrossRef]
  137. Liu, X.; Shao, L.; Liu, X.; Ji, F.; Mei, Y.; Cheng, Y.; Liu, F.; Yan, C.; Li, L.; Ling, Z. Alterations of gastric mucosal microbiota across different stomach microhabitats in a cohort of 276 patients with gastric cancer. EBioMedicine 2019, 40, 336–348. [Google Scholar] [CrossRef] [Green Version]
  138. Hu, Y.-L.; Pang, W.; Huang, Y.; Zhang, Y.; Zhang, C.-J. The gastric microbiome is perturbed in advanced gastric adenocarcinoma identified through shotgun metagenomics. Front. Cell. Infect. Microbiol. 2018, 8, 433. [Google Scholar] [CrossRef] [Green Version]
  139. Aviles-Jimenez, F.; Vazquez-Jimenez, F.; Medrano-Guzman, R.; Mantilla, A.; Torres, J. Stomach microbiota composition varies between patients with non-atrophic gastritis and patients with intestinal type of gastric cancer. Sci. Rep. 2014, 4, 4202. [Google Scholar] [CrossRef] [Green Version]
  140. Sohn, S.-H.; Kim, N.; Jo, H.J.; Kim, J.; Park, J.H.; Nam, R.H.; Seok, Y.-J.; Kim, Y.-R.; Lee, D.H. Analysis of Gastric Body Microbiota by Pyrosequencing: Possible Role of Bacteria Other Than Helicobacter pylori in the Gastric Carcinogenesis. J. Cancer Prev. 2017, 22, 115–125. [Google Scholar] [CrossRef] [Green Version]
  141. Gantuya, B.; El-Serag, H.B.; Matsumoto, T.; Ajami, N.J.; Oyuntsetseg, K.; Azzaya, D.; Uchida, T.; Yamaoka, Y. Gastric Microbiota in Helicobacter pylori-Negative and -Positive Gastritis Among High Incidence of Gastric Cancer Area. Cancers 2019, 11, 504. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  142. Li, Q.; Yu, H. The role of non-H. pylori bacteria in the development of gastric cancer. Am. J. Cancer Res. 2020, 10, 2271–2281. [Google Scholar] [PubMed]
  143. Thorell, K.; Bengtsson-Palme, J.; Liu, O.H.-F.; Gonzales, R.V.P.; Nookaew, I.; Rabeneck, L.; Paszat, L.; Graham, D.Y.; Nielsen, J.; Lundin, S.B.; et al. In Vivo Analysis of the Viable Microbiota and Helicobacter pylori Transcriptome in Gastric Infection and Early Stages of Carcinogenesis. Infect. Immun. 2017, 85, e00031-17. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Molendijk, J.; Nguyen, T.-M.-T.; Brown, I.; Mohamed, A.; Lim, Y.; Barclay, J.; Hodson, M.P.; Hennessy, T.P.; Krause, L.; Morrison, M.; et al. Chronic High-Fat Diet Induces Early Barrett’s Esophagus in Mice through Lipidome Remodeling. Biomolecules 2020, 10, 776. [Google Scholar] [CrossRef]
  145. Hsieh, Y.-Y.; Tung, S.-Y.; Pan, H.-Y.; Yen, C.-W.; Xu, H.-W.; Lin, Y.-J.; Deng, Y.-F.; Hsu, W.-T.; Wu, C.-S.; Li, C. Increased Abundance of Clostridium and Fusobacterium in Gastric Microbiota of Patients with Gastric Cancer in Taiwan. Sci. Rep. 2018, 8, 158. [Google Scholar] [CrossRef]
  146. Boehm, E.T.; Thon, C.; Kupcinskas, J.; Steponaitiene, R.; Skieceviciene, J.; Canbay, A.; Malfertheiner, P.; Link, A. Fusobacterium nucleatum is associated with worse prognosis in Lauren’s diffuse type gastric cancer patients. Sci. Rep. 2020, 10, 16240. [Google Scholar] [CrossRef]
  147. Arita, S.; Ogawa, T.; Murakami, Y.; Kinoshita, Y.; Okazaki, M.; Inagaki-Ohara, K. Dietary Fat-Accelerating Leptin Signaling Promotes Protumorigenic Gastric Environment in Mice. Nutrients 2019, 11, 2127. [Google Scholar] [CrossRef] [Green Version]
  148. Arita, S.; Inagaki-Ohara, K. High-fat-diet–induced modulations of leptin signaling and gastric microbiota drive precancerous lesions in the stomach. Nutrition 2019, 67–68, 110556. [Google Scholar] [CrossRef]
  149. Sears, C.L.; Pardoll, D.M. Perspective: Alpha-bugs, their microbial partners, and the link to colon cancer. J. Infect. Dis. 2011, 203, 306–311. [Google Scholar] [CrossRef] [Green Version]
  150. Tjalsma, H.; Boleij, A.; Marchesi, J.R.; Dutilh, B.E. A bacterial driver-passenger model for colorectal cancer: Beyond the usual suspects. Nat. Rev. Microbiol. 2012, 10, 575–582. [Google Scholar] [CrossRef]
  151. Mori, G.; Rampelli, S.; Orena, B.S.; Rengucci, C.; De Maio, G.; Barbieri, G.; Passardi, A.; Casadei Gardini, A.; Frassineti, G.L.; Gaiarsa, S.; et al. Shifts of Faecal Microbiota During Sporadic Colorectal Carcinogenesis. Sci. Rep. 2018, 8, 10329. [Google Scholar] [CrossRef] [PubMed]
  152. Castellarin, M.; Warren, R.L.; Freeman, J.D.; Dreolini, L.; Krzywinski, M.; Strauss, J.; Barnes, R.; Watson, P.; Allen-Vercoe, E.; Moore, R.A.; et al. Fusobacterium nucleatum infection is prevalent in human colorectal carcinoma. Genome Res. 2012, 22, 299–306. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  153. Yu, T.; Guo, F.; Yu, Y.; Sun, T.; Ma, D.; Han, J.; Qian, Y.; Kryczek, I.; Sun, D.; Nagarsheth, N.; et al. Fusobacterium nucleatum Promotes Chemoresistance to Colorectal Cancer by Modulating Autophagy. Cell 2017, 170, 548–563.e16. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Abed, J.; Emgård, J.E.; Zamir, G.; Faroja, M.; Almogy, G.; Grenov, A.; Sol, A.; Naor, R.; Pikarsky, E.; Atlan, K.A.; et al. Fap2 Mediates Fusobacterium nucleatum Colorectal Adenocarcinoma Enrichment by Binding to Tumor-Expressed Gal-GalNAc. Cell Host Microbe 2016, 20, 215–225. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Bonnet, M.; Buc, E.; Sauvanet, P.; Darcha, C.; Dubois, D.; Pereira, B.; Déchelotte, P.; Bonnet, R.; Pezet, D.; Darfeuille-Michaud, A. Colonization of the Human Gut by E. coli and Colorectal Cancer Risk. Clin. Cancer Res. 2014, 20, 859–867. [Google Scholar] [CrossRef] [Green Version]
  156. Sun, J.; Hobert, M.E.; Duan, Y.; Rao, A.S.; He, T.-C.; Chang, E.B.; Madara, J.L. Crosstalk between NF-κB and β-catenin pathways in bacterial-colonized intestinal epithelial cells. Am. J. Physiol.-Gastrointest. Liver Physiol. 2005, 289, G129–G137. [Google Scholar] [CrossRef]
  157. Moncrief, J.S.; Duncan, A.J.; Wright, R.L.; Barroso, L.A.; Wilkins, T.D. Molecular characterization of the fragilysin pathogenicity islet of enterotoxigenic Bacteroides fragilis. Infect. Immun. 1998, 66, 1735–1739. [Google Scholar] [CrossRef] [Green Version]
  158. Cheng, W.T.; Kantilal, H.K.; Davamani, F. The Mechanism of Bacteroides fragilis Toxin Contributes to Colon Cancer Formation. Malays. J. Med. Sci. 2020, 27, 9–21. [Google Scholar] [CrossRef]
  159. Sears, C.L.; Geis, A.L.; Housseau, F. Bacteroides fragilis subverts mucosal biology: From symbiont to colon carcinogenesis. J. Clin. Investig. 2014, 124, 4166–4172. [Google Scholar] [CrossRef] [Green Version]
  160. Wu, S.; Rhee, K.-J.; Zhang, M.; Franco, A.; Sears, C.L. Bacteroides fragilis toxin stimulates intestinal epithelial cell shedding and γ-secretase-dependent E-cadherin cleavage. J. Cell Sci. 2007, 120, 1944–1952. [Google Scholar] [CrossRef] [Green Version]
  161. Ertz-Archambault, N.; Keim, P.; Von Hoff, D. Microbiome and pancreatic cancer: A comprehensive topic review of literature. World J. Gastroenterol. 2017, 23, 1899–1908. [Google Scholar] [CrossRef]
  162. Michaud, D.S.; Izard, J.; Wilhelm-Benartzi, C.S.; You, D.H.; Grote, V.A.; Tjønneland, A.; Dahm, C.C.; Overvad, K.; Jenab, M.; Fedirko, V.; et al. Plasma antibodies to oral bacteria and risk of pancreatic cancer in a large European prospective cohort study. Gut 2013, 62, 1764–1770. [Google Scholar] [CrossRef] [PubMed]
  163. Fan, X.; Alekseyenko, A.V.; Wu, J.; Peters, B.A.; Jacobs, E.J.; Gapstur, S.M.; Purdue, M.P.; Abnet, C.C.; Stolzenberg-Solomon, R.; Miller, G.; et al. Human oral microbiome and prospective risk for pancreatic cancer: A population-based nested case-control study. Gut 2018, 67, 120–127. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  164. Wei, M.Y.; Shi, S.; Liang, C.; Meng, Q.C.; Hua, J.; Zhang, Y.Y.; Liu, J.; Zhang, B.; Xu, J.; Yu, X.J. The microbiota and microbiome in pancreatic cancer: More influential than expected. Mol. Cancer 2019, 18, 97. [Google Scholar] [CrossRef] [PubMed]
  165. Jesnowski, R.; Isaksson, B.; Möhrcke, C.; Bertsch, C.; Bulajic, M.; Schneider-Brachert, W.; Klöppel, G.; Lowenfels, A.B.; Maisonneuve, P.; Löhr, J.M. Helicobacter pylori in autoimmune pancreatitis and pancreatic carcinoma. Pancreatology 2010, 10, 462–466. [Google Scholar] [CrossRef]
  166. Knorr, J.; Ricci, V.; Hatakeyama, M.; Backert, S. Classification of Helicobacter pylori Virulence Factors: Is CagA a Toxin or Not? Trends Microbiol. 2019, 27, 731–738. [Google Scholar] [CrossRef] [PubMed]
  167. Pushalkar, S.; Hundeyin, M.; Daley, D.; Zambirinis, C.P.; Kurz, E.; Mishra, A.; Mohan, N.; Aykut, B.; Usyk, M.; Torres, L.E.; et al. The Pancreatic Cancer Microbiome Promotes Oncogenesis by Induction of Innate and Adaptive Immune Suppression. Cancer Discov. 2018, 8, 403–416. [Google Scholar] [CrossRef] [Green Version]
  168. Thomas, R.M.; Gharaibeh, R.Z.; Gauthier, J.; Beveridge, M.; Pope, J.L.; Guijarro, M.V.; Yu, Q.; He, Z.; Ohland, C.; Newsome, R.; et al. Intestinal microbiota enhances pancreatic carcinogenesis in preclinical models. Carcinogenesis 2018, 39, 1068–1078. [Google Scholar] [CrossRef]
  169. Benner, M.; Ferwerda, G.; Joosten, I.; van der Molen, R.G. How uterine microbiota might be responsible for a receptive, fertile endometrium. Hum. Reprod. Update 2018, 24, 393–415. [Google Scholar] [CrossRef] [Green Version]
  170. Chen, C.; Song, X.; Wei, W.; Zhong, H.; Dai, J.; Lan, Z.; Li, F.; Yu, X.; Feng, Q.; Wang, Z.; et al. The microbiota continuum along the female reproductive tract and its relation to uterine-related diseases. Nat. Commun. 2017, 8, 875. [Google Scholar] [CrossRef] [Green Version]
  171. Ravel, J.; Gajer, P.; Abdo, Z.; Schneider, G.M.; Koenig, S.S.; McCulle, S.L.; Karlebach, S.; Gorle, R.; Russell, J.; Tacket, C.O.; et al. Vaginal microbiome of reproductive-age women. Proc. Natl. Acad. Sci. USA 2011, 108 (Suppl. S1), 4680–4687. [Google Scholar] [CrossRef] [Green Version]
  172. Smith, B.C.; Zolnik, C.P.; Usyk, M.; Chen, Z.; Kaiser, K.; Nucci-Sack, A.; Peake, K.; Diaz, A.; Viswanathan, S.; Strickler, H.D.; et al. Distinct Ecological Niche of Anal, Oral, and Cervical Mucosal Microbiomes in Adolescent Women. Yale J. Biol. Med. 2016, 89, 277–284. [Google Scholar]
  173. Dethlefsen, L.; McFall-Ngai, M.; Relman, D.A. An ecological and evolutionary perspective on human-microbe mutualism and disease. Nature 2007, 449, 811–818. [Google Scholar] [CrossRef] [PubMed]
  174. Valenti, P.; Rosa, L.; Capobianco, D.; Lepanto, M.S.; Schiavi, E.; Cutone, A.; Paesano, R.; Mastromarino, P. Role of Lactobacilli and Lactoferrin in the Mucosal Cervicovaginal Defense. Front. Immunol. 2018, 9, 376. [Google Scholar] [CrossRef] [PubMed]
  175. Petrova, M.I.; Lievens, E.; Malik, S.; Imholz, N.; Lebeer, S. Lactobacillus species as biomarkers and agents that can promote various aspects of vaginal health. Front. Physiol. 2015, 6, 81. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  176. Ma, B.; Forney, L.J.; Ravel, J. Vaginal microbiome: Rethinking health and disease. Annu. Rev. Microbiol. 2012, 66, 371–389. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  177. Smith, S.B.; Ravel, J. The vaginal microbiota, host defence and reproductive physiology. J. Physiol. 2017, 595, 451–463. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  178. Bik, E.M.; Bird, S.W.; Bustamante, J.P.; Leon, L.E.; Nieto, P.A.; Addae, K.; Alegría-Mera, V.; Bravo, C.; Bravo, D.; Cardenas, J.P.; et al. A novel sequencing-based vaginal health assay combining self-sampling, HPV detection and genotyping, STI detection, and vaginal microbiome analysis. PLoS ONE 2019, 14, e0215945. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  179. Freitas, A.C.; Bocking, A.; Hill, J.E.; Money, D.M. Increased richness and diversity of the vaginal microbiota and spontaneous preterm birth. Microbiome 2018, 6, 117. [Google Scholar] [CrossRef] [PubMed]
  180. Huang, B.; Fettweis, J.M.; Brooks, J.P.; Jefferson, K.K.; Buck, G.A. The changing landscape of the vaginal microbiome. Clin. Lab. Med. 2014, 34, 747–761. [Google Scholar] [CrossRef] [Green Version]
  181. Torcia, M.G. Interplay among Vaginal Microbiome, Immune Response and Sexually Transmitted Viral Infections. Int. J. Mol. Sci. 2019, 20, 266. [Google Scholar] [CrossRef] [Green Version]
  182. Garrett, W.S. Cancer and the microbiota. Science 2015, 348, 80–86. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  183. Schwabe, R.F.; Jobin, C. The microbiome and cancer. Nat. Rev. Cancer 2013, 13, 800–812. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  184. Walther-António, M.R.; Chen, J.; Multinu, F.; Hokenstad, A.; Distad, T.J.; Cheek, E.H.; Keeney, G.L.; Creedon, D.J.; Nelson, H.; Mariani, A.; et al. Potential contribution of the uterine microbiome in the development of endometrial cancer. Genome Med. 2016, 8, 122. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  185. Mitra, A.; MacIntyre, D.A.; Lee, Y.S.; Smith, A.; Marchesi, J.R.; Lehne, B.; Bhatia, R.; Lyons, D.; Paraskevaidis, E.; Li, J.V.; et al. Cervical intraepithelial neoplasia disease progression is associated with increased vaginal microbiome diversity. Sci. Rep. 2015, 5, 16865. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  186. Di Paola, M.; Sani, C.; Clemente, A.M.; Iossa, A.; Perissi, E.; Castronovo, G.; Tanturli, M.; Rivero, D.; Cozzolino, F.; Cavalieri, D.; et al. Characterization of cervico-vaginal microbiota in women developing persistent high-risk Human Papillomavirus infection. Sci. Rep. 2017, 7, 10200. [Google Scholar] [CrossRef]
  187. Lee, J.E.; Lee, S.; Lee, H.; Song, Y.M.; Lee, K.; Han, M.J.; Sung, J.; Ko, G. Association of the vaginal microbiota with human papillomavirus infection in a Korean twin cohort. PLoS ONE 2013, 8, e63514. [Google Scholar] [CrossRef] [PubMed]
  188. Norenhag, J.; Du, J.; Olovsson, M.; Verstraelen, H.; Engstrand, L.; Brusselaers, N. The vaginal microbiota, human papillomavirus and cervical dysplasia: A systematic review and network meta-analysis. BJOG 2020, 127, 171–180. [Google Scholar] [CrossRef]
  189. Kwasniewski, W.; Wolun-Cholewa, M.; Kotarski, J.; Warchol, W.; Kuzma, D.; Kwasniewska, A.; Gozdzicka-Jozefiak, A. Microbiota dysbiosis is associated with HPV-induced cervical carcinogenesis. Oncol. Lett. 2018, 16, 7035–7047. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  190. Brusselaers, N.; Shrestha, S.; van de Wijgert, J.; Verstraelen, H. Vaginal dysbiosis and the risk of human papillomavirus and cervical cancer: Systematic review and meta-analysis. Am. J. Obstet. Gynecol. 2019, 221, 9–18.e18. [Google Scholar] [CrossRef]
  191. Kyrgiou, M.; Moscicki, A.B. Vaginal microbiome and cervical cancer. Semin. Cancer Biol. 2022; in press. [Google Scholar] [CrossRef] [PubMed]
  192. Baker, J.M.; Al-Nakkash, L.; Herbst-Kralovetz, M.M. Estrogen–gut microbiome axis: Physiological and clinical implications. Maturitas 2017, 103, 45–53. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  193. Ding, L.; Liu, C.; Zhou, Q.; Feng, M.; Wang, J. Association of estradiol and HPV/HPV16 infection with the occurrence of cervical squamous cell carcinoma. Oncol. Lett. 2019, 17, 3548–3554. [Google Scholar] [CrossRef] [Green Version]
  194. James, C.D.; Morgan, I.M.; Bristol, M.L. The relationship between estrogen-related signaling and human papillomavirus positive cancers. Pathogens 2020, 9, 403. [Google Scholar] [CrossRef] [PubMed]
  195. Yuan, X.-Y.; Liu, H.-Z.; Liu, J.-F.; Sun, Y.; Song, Y. Pathogenic mechanism, detection methods and clinical significance of group B Streptococcus. Future Microbiol. 2021, 16, 671–685. [Google Scholar] [CrossRef] [PubMed]
  196. Antonio, M.A.; Rabe, L.K.; Hillier, S.L. Colonization of the rectum by Lactobacillus species and decreased risk of bacterial vaginosis. J. Infect. Dis. 2005, 192, 394–398. [Google Scholar] [CrossRef] [Green Version]
  197. Homayouni, A.; Bastani, P.; Ziyadi, S.; Mohammad-Alizadeh-Charandabi, S.; Ghalibaf, M.; Mortazavian, A.M.; Mehrabany, E.V. Effects of probiotics on the recurrence of bacterial vaginosis: A review. J. Low. Genit. Tract Dis. 2014, 18, 79–86. [Google Scholar] [CrossRef]
  198. Blaak, E.; Canfora, E.; Theis, S.; Frost, G.; Groen, A.; Mithieux, G.; Nauta, A.; Scott, K.; Stahl, B.; van Harsselaar, J. Short chain fatty acids in human gut and metabolic health. Benef. Microbes 2020, 11, 411–455. [Google Scholar] [CrossRef] [PubMed]
  199. Delgado-Diaz, D.J.; Tyssen, D.; Hayward, J.A.; Gugasyan, R.; Hearps, A.C.; Tachedjian, G. Distinct immune responses elicited from cervicovaginal epithelial cells by lactic acid and short chain fatty acids associated with optimal and non-optimal vaginal microbiota. Front. Cell. Infect. Microbiol. 2020, 9, 446. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  200. Xu, S.; Liu, Z.; Lv, M.; Chen, Y.; Liu, Y. Intestinal dysbiosis promotes epithelial-mesenchymal transition by activating tumor-associated macrophages in ovarian cancer. Pathog. Dis 2019, 77, ftz019. [Google Scholar] [CrossRef]
  201. Magnus, M.C.; Håberg, S.E.; Stigum, H.; Nafstad, P.; London, S.J.; Vangen, S.; Nystad, W. Delivery by Cesarean section and early childhood respiratory symptoms and disorders: The Norwegian mother and child cohort study. Am. J. Epidemiol. 2011, 174, 1275–1285. [Google Scholar] [CrossRef]
  202. Almqvist, C.; Cnattingius, S.; Lichtenstein, P.; Lundholm, C. The impact of birth mode of delivery on childhood asthma and allergic diseases—A sibling study. Clin. Exp. Allergy 2012, 42, 1369–1376. [Google Scholar] [CrossRef]
  203. Gondwe, T.; Betha, K.; Kusneniwar, G.; Bunker, C.H.; Tang, G.; Simhan, H.; Reddy, P.; Haggerty, C.L. Mode of delivery and short-term infant health outcomes: A prospective cohort study in a peri-urban Indian population. BMC Pediatr. 2018, 18, 346. [Google Scholar] [CrossRef] [Green Version]
  204. Mima, K.; Nishihara, R.; Qian, Z.R.; Cao, Y.; Sukawa, Y.; Nowak, J.A.; Yang, J.; Dou, R.; Masugi, Y.; Song, M.; et al. Fusobacterium nucleatum in colorectal carcinoma tissue and patient prognosis. Gut 2016, 65, 1973–1980. [Google Scholar] [CrossRef] [Green Version]
  205. Mira-Pascual, L.; Cabrera-Rubio, R.; Ocon, S.; Costales, P.; Parra, A.; Suarez, A.; Moris, F.; Rodrigo, L.; Mira, A.; Collado, M.C. Microbial mucosal colonic shifts associated with the development of colorectal cancer reveal the presence of different bacterial and archaeal biomarkers. J. Gastroenterol. 2015, 50, 167–179. [Google Scholar] [CrossRef]
  206. Tahara, T.; Yamamoto, E.; Suzuki, H.; Maruyama, R.; Chung, W.; Garriga, J.; Jelinek, J.; Yamano, H.-O.; Sugai, T.; An, B.; et al. Fusobacterium in colonic flora and molecular features of colorectal carcinoma. Cancer Res. 2014, 74, 1311–1318. [Google Scholar] [CrossRef] [Green Version]
  207. Nugent, J.L.; McCoy, A.N.; Addamo, C.J.; Jia, W.; Sandler, R.S.; Keku, T.O. Altered tissue metabolites correlate with microbial dysbiosis in colorectal adenomas. J. Proteome Res. 2014, 13, 1921–1929. [Google Scholar] [CrossRef]
  208. Wu, N.; Yang, X.; Zhang, R.; Li, J.; Xiao, X.; Hu, Y.; Chen, Y.; Yang, F.; Lu, N.; Wang, Z.; et al. Dysbiosis signature of fecal microbiota in colorectal.l cancer patients. Microb. Ecol. 2013, 66, 462–470. [Google Scholar] [CrossRef]
  209. Warren, R.L.; Freeman, D.J.; Pleasance, S.; Watson, P.; Moore, R.A.; Cochrane, K.; Allen-Vercoe, E.; Holt, R.A. Co-occurrence of anaerobic bacteria in colorectal carcinomas. Microbiome 2013, 1, 16. [Google Scholar] [CrossRef] [Green Version]
  210. McCoy, A.N.; Araújo-Pérez, F.; Azcárate-Peril, A.; Yeh, J.J.; Sandler, R.S.; Keku, T.O. Fusobacterium is associated with colorectal adenomas. PLoS ONE 2013, 8, e53653. [Google Scholar] [CrossRef] [PubMed]
  211. Brim, H.; Yooseph, S.; Zoetendal, E.G.; Lee, E.; Torralbo, M.; Laiyemo, A.O.; Shokrani, B.; Nelson, K.; Ashktorab, H. Microbiome analysis of stool samples from African Americans with colon polyps. PLoS ONE 2013, 8, e81352. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  212. Sanapareddy, N.; Legge, R.M.; Jovov, B.; McCoy, A.; Burcal, L.; Araujo-Perez, F.; Randall, T.A.; Galanko, J.; Benson, A.; Sandler, R.S.; et al. Increased rectal microbial richness is associated with the presence of colorectal adenomas in humans. ISME J. 2012, 6, 1858–1868. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  213. Marchesi, J.R.; Dutilh, B.E.; Hall, N.; Peters, W.H.M.; Roelofs, R.; Boleij, A.; Tjalsma, H. Towards the Human Colorectal Cancer Microbiome. PLoS ONE 2011, 6, e20447. [Google Scholar] [CrossRef] [Green Version]
  214. Shen, X.J.; Rawls, J.F.; Randall, T.; Burcal, L.; Mpande, C.N.; Jenkins, N.; Jovov, B.; Abdo, Z.; Sandler, R.S.; Keku, T.O. Molecular characterization of mucosal adherent bacteria and associations with colorectal adenomas. Gut Microbes 2010, 1, 138–147. [Google Scholar] [CrossRef] [Green Version]
  215. Dicksved, J.; Lindberg, M.; Rosenquist, M.; Enroth, H.; Jansson, J.K.; Engstrand, L. Molecular characterization of the stomach microbiota in patients with gastric cancer and in controls. J. Med. Microbiol. 2009, 58, 509–516. [Google Scholar] [CrossRef]
  216. Sobhani, I.; Tap, J.; Roudot-Thoraval, F.; Roperch, J.P.; Letulle, S.; Langella, P.; Corthier, G.; Tran Van Nhieu, J.; Furet, J.P. Microbial dysbiosis in colorectal cancer (CRC) patients. PLoS ONE 2011, 6, e16393. [Google Scholar] [CrossRef]
  217. Apostolou, P.; Tsantsaridou, A.; Papasotiriou, I.; Toloudi, M.; Chatziioannou, M.; Giamouzis, G. Bacterial and fungal microflora in surgically removed lung cancer samples. J. Cardiothorac. Surg. 2011, 6, 137. [Google Scholar] [CrossRef] [Green Version]
  218. Chan, P.J.; Seraj, I.M.; Kalugdan, T.H.; King, A. Prevalence of Mycoplasma Conserved DNA in Malignant Ovarian Cancer Detected Using Sensitive PCR–ELISA. Gynecol. Oncol. 1996, 63, 258–260. [Google Scholar] [CrossRef]
  219. van Vliet, M.J.; Tissing, W.J.; Dun, C.A.; Meessen, N.E.; Kamps, W.A.; de Bont, E.S.; Harmsen, H.J. Chemotherapy treatment in pediatric patients with acute myeloid leukemia receiving antimicrobial prophylaxis leads to a relative increase of colonization with potentially pathogenic bacteria in the gut. Clin. Infect. Dis. 2009, 49, 262–270. [Google Scholar] [CrossRef] [Green Version]
  220. Jenq, R.R.; Ubeda, C.; Taur, Y.; Menezes, C.C.; Khanin, R.; Dudakov, J.A.; Liu, C.; West, M.L.; Singer, N.V.; Equinda, M.J.; et al. Regulation of intestinal inflammation by microbiota following allogeneic bone marrow transplantation. J. Exp. Med. 2012, 209, 903–911. [Google Scholar] [CrossRef]
  221. Viaud, S.; Saccheri, F.; Mignot, G.; Yamazaki, T.; Daillère, R.; Hannani, D.; Enot, D.P.; Pfirschke, C.; Engblom, C.; Pittet, M.J.; et al. The intestinal microbiota modulates the anticancer immune effects of cyclophosphamide. Science 2013, 342, 971–976. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  222. Iida, N.; Dzutsev, A.; Stewart, C.A.; Smith, L.; Bouladoux, N.; Weingarten, R.A.; Molina, D.A.; Salcedo, R.; Back, T.; Cramer, S.; et al. Commensal bacteria control cancer response to therapy by modulating the tumor microenvironment. Science 2013, 342, 967–970. [Google Scholar] [CrossRef] [PubMed]
  223. Alexander, J.L.; Wilson, I.D.; Teare, J.; Marchesi, J.R.; Nicholson, J.K.; Kinross, J.M. Gut microbiota modulation of chemotherapy efficacy and toxicity. Nat. Rev. Gastroenterol. Hepatol. 2017, 14, 356–365. [Google Scholar] [CrossRef]
  224. Montassier, E.; Gastinne, T.; Vangay, P.; Al-Ghalith, G.A.; Bruley des Varannes, S.; Massart, S.; Moreau, P.; Potel, G.; de La Cochetière, M.F.; Batard, E.; et al. Chemotherapy-driven dysbiosis in the intestinal microbiome. Aliment. Pharmacol. Ther. 2015, 42, 515–528. [Google Scholar] [CrossRef]
  225. Gerassy-Vainberg, S.; Blatt, A.; Danin-Poleg, Y.; Gershovich, K.; Sabo, E.; Nevelsky, A.; Daniel, S.; Dahan, A.; Ziv, O.; Dheer, R.; et al. Radiation induces proinflammatory dysbiosis: Transmission of inflammatory susceptibility by host cytokine induction. Gut 2018, 67, 97–107. [Google Scholar] [CrossRef] [Green Version]
  226. Schwan, A.; Sjölin, S.; Trottestam, U.; Aronsson, B. Relapsing clostridium difficile enterocolitis cured by rectal infusion of homologous faeces. Lancet 1983, 2, 845. [Google Scholar] [CrossRef]
  227. Gough, E.; Shaikh, H.; Manges, A.R. Systematic review of intestinal microbiota transplantation (fecal bacteriotherapy) for recurrent Clostridium difficile infection. Clin. Infect. Dis. 2011, 53, 994–1002. [Google Scholar] [CrossRef] [Green Version]
  228. Brandt, L.J. Fecal Microbiota Transplant: Respice, Adspice, Prospice. J. Clin. Gastroenterol. 2015, 49 (Suppl. S1), S65–S68. [Google Scholar] [CrossRef]
  229. Kao, D.; Roach, B.; Silva, M.; Beck, P.; Rioux, K.; Kaplan, G.G.; Chang, H.-J.; Coward, S.; Goodman, K.J.; Xu, H.; et al. Effect of Oral Capsule– vs Colonoscopy-Delivered Fecal Microbiota Transplantation on Recurrent Clostridium difficile Infection: A Randomized Clinical Trial. JAMA 2017, 318, 1985–1993. [Google Scholar] [CrossRef] [Green Version]
  230. Khanna, S.; Raffals, L.E. The Microbiome in Crohn’s Disease: Role in Pathogenesis and Role of Microbiome Replacement Therapies. Gastroenterol. Clin. N. Am. 2017, 46, 481–492. [Google Scholar] [CrossRef]
  231. Khanna, S. Microbiota Replacement Therapies: Innovation in Gastrointestinal Care. Clin. Pharmacol. Ther. 2018, 103, 102–111. [Google Scholar] [CrossRef] [PubMed]
  232. Vindigni, S.M.; Surawicz, C.M. Fecal Microbiota Transplantation. Gastroenterol. Clin. N. Am. 2017, 46, 171–185. [Google Scholar] [CrossRef] [PubMed]
  233. Gupta, A.; Saha, S.; Khanna, S. Therapies to modulate gut microbiota: Past, present and future. World J. Gastroenterol. 2020, 26, 777–788. [Google Scholar] [CrossRef] [PubMed]
  234. Davar, D.; Dzutsev, A.K.; McCulloch, J.A.; Rodrigues, R.R.; Chauvin, J.M.; Morrison, R.M.; Deblasio, R.N.; Menna, C.; Ding, Q.; Pagliano, O.; et al. Fecal microbiota transplant overcomes resistance to anti-PD-1 therapy in melanoma patients. Science 2021, 371, 595–602. [Google Scholar] [CrossRef]
  235. Mohty, M.; Malard, F.; D’Incan, E.; Thomas, X.; Recher, C.; Michallet, A.-S.; Peterlin, P.; Vekhoff, A.; Vey, N.; Plantamura, E. Prevention of dysbiosis complications with autologous fecal microbiota transplantation (auto-FMT) in acute myeloid leukemia (AML) patients undergoing intensive treatment (ODYSSEE study): First results of a prospective multicenter trial. Blood 2017, 130, 2624. [Google Scholar]
  236. Taur, Y.; Jenq, R.R.; Ubeda, C.; van den Brink, M.; Pamer, E.G. Role of intestinal microbiota in transplantation outcomes. Best Pract. Res. Clin. Haematol. 2015, 28, 155–161. [Google Scholar] [CrossRef] [Green Version]
  237. Guarner, F.; Schaafsma, G.J. Probiotics. Int. J. Food Microbiol. 1998, 39, 237–238. [Google Scholar] [CrossRef]
  238. Zhu, Y.; Michelle Luo, T.; Jobin, C.; Young, H.A. Gut microbiota and probiotics in colon tumorigenesis. Cancer Lett. 2011, 309, 119–127. [Google Scholar] [CrossRef] [Green Version]
  239. Appleyard, C.B.; Cruz, M.L.; Isidro, A.A.; Arthur, J.C.; Jobin, C.; De Simone, C. Pretreatment with the probiotic VSL#3 delays transition from inflammation to dysplasia in a rat model of colitis-associated cancer. Am. J. Physiol.-Gastrointest. Liver Physiol. 2011, 301, G1004–G1013. [Google Scholar] [CrossRef] [Green Version]
  240. Arthur, J.C.; Gharaibeh, R.Z.; Uronis, J.M.; Perez-Chanona, E.; Sha, W.; Tomkovich, S.; Mühlbauer, M.; Fodor, A.A.; Jobin, C. VSL#3 probiotic modifies mucosal microbial composition but does not reduce colitis-associated colorectal cancer. Sci. Rep. 2013, 3, 2868. [Google Scholar] [CrossRef] [Green Version]
  241. Hibberd, A.A.; Lyra, A.; Ouwehand, A.C.; Rolny, P.; Lindegren, H.; Cedgård, L.; Wettergren, Y. Intestinal microbiota is altered in patients with colon cancer and modified by probiotic intervention. BMJ Open Gastroenterol. 2017, 4, e000145. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  242. Gianotti, L.; Morelli, L.; Galbiati, F.; Rocchetti, S.; Coppola, S.; Beneduce, A.; Gilardini, C.; Zonenschain, D.; Nespoli, A.; Braga, M. A randomized double-blind trial on perioperative administration of probiotics in colorectal cancer patients. World J. Gastroenterol. 2010, 16, 167–175. [Google Scholar] [CrossRef] [PubMed]
  243. Riehl, T.E.; Alvarado, D.; Ee, X.; Zuckerman, A.; Foster, L.; Kapoor, V.; Thotala, D.; Ciorba, M.A.; Stenson, W.F. Lactobacillus rhamnosus GG protects the intestinal epithelium from radiation injury through release of lipoteichoic acid, macrophage activation and the migration of mesenchymal stem cells. Gut 2019, 68, 1003–1013. [Google Scholar] [CrossRef] [PubMed]
  244. Dizman, N.; Hsu, J.; Bergerot, P.G.; Gillece, J.D.; Folkerts, M.; Reining, L.; Trent, J.; Highlander, S.K.; Pal, S.K. Randomized trial assessing impact of probiotic supplementation on gut microbiome and clinical outcome from targeted therapy in metastatic renal cell carcinoma. Cancer Med. 2021, 10, 79–86. [Google Scholar] [CrossRef] [PubMed]
  245. Tian, Y.; Li, M.; Song, W.; Jiang, R.; Li, Y.Q. Effects of probiotics on chemotherapy in patients with lung cancer. Oncol. Lett. 2019, 17, 2836–2848. [Google Scholar] [CrossRef] [Green Version]
  246. Carmody, R.N.; Gerber, G.K.; Luevano, J.M., Jr.; Gatti, D.M.; Somes, L.; Svenson, K.L.; Turnbaugh, P.J. Diet dominates host genotype in shaping the murine gut microbiota. Cell Host Microbe 2015, 17, 72–84. [Google Scholar] [CrossRef] [Green Version]
  247. Le Chatelier, E.; Nielsen, T.; Qin, J.; Prifti, E.; Hildebrand, F.; Falony, G.; Almeida, M.; Arumugam, M.; Batto, J.-M.; Kennedy, S.; et al. Richness of human gut microbiome correlates with metabolic markers. Nature 2013, 500, 541–546. [Google Scholar] [CrossRef]
  248. David, L.A.; Maurice, C.F.; Carmody, R.N.; Gootenberg, D.B.; Button, J.E.; Wolfe, B.E.; Ling, A.V.; Devlin, A.S.; Varma, Y.; Fischbach, M.A.; et al. Diet rapidly and reproducibly alters the human gut microbiome. Nature 2014, 505, 559–563. [Google Scholar] [CrossRef] [Green Version]
  249. Cotillard, A.; Kennedy, S.P.; Kong, L.C.; Prifti, E.; Pons, N.; Le Chatelier, E.; Almeida, M.; Quinquis, B.; Levenez, F.; Galleron, N.; et al. Dietary intervention impact on gut microbial gene richness. Nature 2013, 500, 585–588. [Google Scholar] [CrossRef]
  250. Taper, H.S.; Roberfroid, M.B. Possible adjuvant cancer therapy by two prebiotics--inulin or oligofructose. In Vivo 2005, 19, 201–204. [Google Scholar]
  251. O’Keefe, S.J.D. Diet, microorganisms and their metabolites, and colon cancer. Nat. Rev. Gastroenterol. Hepatol. 2016, 13, 691–706. [Google Scholar] [CrossRef] [PubMed]
  252. Jakobsson, H.E.; Jernberg, C.; Andersson, A.F.; Sjölund-Karlsson, M.; Jansson, J.K.; Engstrand, L. Short-term antibiotic treatment has differing long-term impacts on the human throat and gut microbiome. PLoS ONE 2010, 5, e9836. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  253. Pflug, N.; Kluth, S.; Vehreschild, J.J.; Bahlo, J.; Tacke, D.; Biehl, L.; Eichhorst, B.; Fischer, K.; Cramer, P.; Fink, A.M.; et al. Efficacy of antineoplastic treatment is associated with the use of antibiotics that modulate intestinal microbiota. Oncoimmunology 2016, 5, e1150399. [Google Scholar] [CrossRef] [PubMed]
  254. Cieplak, T.; Soffer, N.; Sulakvelidze, A.; Nielsen, D.S. A bacteriophage cocktail targeting Escherichia coli reduces E. coli in simulated gut conditions, while preserving a non-targeted representative commensal normal microbiota. Gut Microbes 2018, 9, 391–399. [Google Scholar] [CrossRef] [Green Version]
  255. Zuo, T.; Wong, S.H.; Lam, K.; Lui, R.; Cheung, K.; Tang, W.; Ching, J.Y.L.; Chan, P.K.S.; Chan, M.C.W.; Wu, J.C.Y.; et al. Bacteriophage transfer during faecal microbiota transplantation in Clostridium difficile infection is associated with treatment outcome. Gut 2018, 67, 634–643. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. Molecular pathways depicting the microbiome-induced EMT and chronic inflammation. F. nucleatum: E. coli strains producing genotoxic compound colibactin can bind to the DNA leading to DNA damage by triggering reactive oxygen species (ROS) and activating the Erk pathway. Activation of Erk stimulates Vimentin and N-cadherin expression, leading to EMT. Microbes express microorganism-associated molecular patterns (MAMPs) and are recognized by macrophages via TLRs. They can either produce ROS from macrophages or trigger the production of proinflammatory cytokines (IL-1, IL-6, IL-8, IL-23, and TNF) via various signaling pathways. Proinflammatory cytokines can activate STAT3 and NF-κB signaling, leading to activation of c-myc oncogene and MMP13, respectively, which progress to EMT, chronic inflammation, and eventually cancer. Simultaneously, virulence factors, FadA and BFT, can disrupt E-cadherin and trigger β-catenin/Wnt signaling pathways resulting in subsequent activation of the STAT3 and NF-κB pathways.
Figure 1. Molecular pathways depicting the microbiome-induced EMT and chronic inflammation. F. nucleatum: E. coli strains producing genotoxic compound colibactin can bind to the DNA leading to DNA damage by triggering reactive oxygen species (ROS) and activating the Erk pathway. Activation of Erk stimulates Vimentin and N-cadherin expression, leading to EMT. Microbes express microorganism-associated molecular patterns (MAMPs) and are recognized by macrophages via TLRs. They can either produce ROS from macrophages or trigger the production of proinflammatory cytokines (IL-1, IL-6, IL-8, IL-23, and TNF) via various signaling pathways. Proinflammatory cytokines can activate STAT3 and NF-κB signaling, leading to activation of c-myc oncogene and MMP13, respectively, which progress to EMT, chronic inflammation, and eventually cancer. Simultaneously, virulence factors, FadA and BFT, can disrupt E-cadherin and trigger β-catenin/Wnt signaling pathways resulting in subsequent activation of the STAT3 and NF-κB pathways.
Cancers 14 02964 g001
Table 1. Overview of the studies exploring the gut-microbiota-associated human cancers.
Table 1. Overview of the studies exploring the gut-microbiota-associated human cancers.
StudyDetection MethodBacterium SpeciesExpression Levels
Colorectal Cancer
Boehm et al. (2020) [146]Probe-based quantitative PCRFusobacterium nucleatumUpregulated
Mori et al. (2018) [151]16S rRNA gene sequencingSutterella and Escherichia/ShigellaUpregulated
Yu et al. (2017) [153]Quantitative PCRFusobacterium nucleatumUpregulated
Mima et al. (2015) [204]Molecular pathological epidemiology databaseFusobacterium nucleatumUpregulated
Mira-Pascual et al. (2015) [205]16S rRNA gene pyrosequencing and quantitative PCRMethanobacteriales, Methanobrevibacterium, Fusobacterium nucleatum, Enterobacteriaceae, Akkermansia muciniphila, and Blautia coccoidesUpregulated
Bifidobacterium, Faecalibacterium prausnitzii, and LactobacillusDownregulated
Tahara et al. (2014) [206]Quantitative real-time PCRFusobacterium nucleatum and pan-fusobacteriumUpregulated
Zackular et al. (2014) [32] 16S rRNA gene sequencingRuminococcaceae, Clostridium, Pseudomonas, and PorphyromonadaceaeUpregulated
Bonnet et al. (2014) [155]PCREscherichia coliUpregulated
Nugent et al. (2014) [207]Quantitative real-time PCRBifidobacterium, Eubacteria, Escherichia coli, Clostridium, and BacteroidesUpregulated
Wu et al. (2013) [208]Pyrosequencing of the 16S rRNA gene V3 regionBacteroids, Fusobacterium, and CampylobacterUpregulated
Faecalibacterium and RoseburiaDownregulated
Warren et al. (2013) [209]Metatranscriptomic analysisFusobacterium, Leptotrichia, and CampylobacterUpregulated
McCoy et al. (2013) [210]16S rRNA quantitative PCR and pyrosequencingFusobacteriumUpregulated
Brim et al. (2013) [211]Human intestinal Tract Chip (HITChip) and 16S rRNA gene barcoded 454 pyrosequencingBacteroidetes and FirmicutesUpregulated
Castellarin et al. (2012) [152]Quantitative PCRFusobacterium nucleatumUpregulated
Sanapareddy et al. (2012) [212]454 titanium pyrosequencing of the V1–V2 region of the 16S rRNA geneFirmicutes, Bacteroidetes, Pseudomonas, Helicobacter, Actinobacteria, Lactobacillus, Acinetobacter, and ProteobacteriaUpregulated
Marchesi et al. (2011) [213]Deep rRNA sequencingRoseburia, Fusobacterium, and FaecalibacteriumUpregulated
Citrobacter, Shigella, Cronobacter, Kluyvera, Serratia, and Salmonella spp.Downregulated
Shen et al. (2010) [214]Terminal restriction fragment length polymorphism, clone sequencing and fluorescent in situ hybridization analysis of the 16S rRNA genesDorea spp. and Faecalibacterium spp.Upregulated
Esophageal Cancer
Nie et al. (2014) [128]Meta-analysisHelicobacter pyloriDownregulated
Chow et al. (1998) [127]Antigen-specific ELISAHelicobacter pyloriDownregulated
Gastric Cancer
Boehm et al. (2020) [146]Probe-based quantitative PCRFusobacterium nucleatumUpregulated
Hansen et al. (2020) [134]18S rDNA sequencingMalasseziaUpregulated
Hsieh et al. (2018) [145]16S ribosomal DNA analysisFusobacterium and ClostridiumUpregulated
Helicobacter pyloriDownregulated
Ferriera et al. (2018) [132]16S rRNA next-generation sequencingHelicobacter pyloriDownregulated
Yu et al. (2017) [136]16S rRNA gene sequencingHelicobacter pyloriUpregulated
Sohn et al. (2017) [140]Bar-coded 454 pyrosequencing of the 16S rRNA geneStreptococcus pseudopneumoniae, S. parasanguinis, and S. oralisUpregulated
Aviles-Jimenez et al. (2014) [139]Microarray G3 PhyloChip analysisPseudomonas, Lactobacillus coleohominis, and LachnospiraceaeUpregulated
Porphyromonas, TM7, Neisseria, and Streptococcus sinensisDownregulated
Dicksved et al. (2009) [215]Terminal restriction fragment length polymorphism analysis in combination with 16S rRNA gene cloning and sequencingStreptococcus, Lactobacillus, Veillonella, and PrevotellaUpregulated
Chow et al. (1998) [127]Antigen-specific ELISAHelicobacter pyloriDownregulated
Lung Cancer
Sobhani et al. (2011) [216]Quantitative PCR and pyrosequencingHelicobacter pyloriDownregulated
Bifidobacterium, Faecalibacterium, Streptococcus, and VeillonellaDownregulated
Gui et al. (2020) [84]Quantitative PCRFaecalibacterium prausnitzii, Clostridium leptum, Ruminococcus spp., Clostridial cluster I, Clostridial cluster XIVa, and Roseburia spp.Downregulated
Zhuang at el. (2019) [82]16S rRNA next-generation sequencingEnterococcusUpregulated
BifidobacteriumDownregulated
Liu et al. (2019) [137]16S rRNA gene amplicon sequencingFusobacteria, Prevotella
Proteobacteria, Streptococcus, Verrucomicrobia, and Veillonella
Upregulated
Bacteroidetes, Firmicutes, and ActinobacteriaDownregulated
Zhang et al. (2018) [37]16S rRNA gene sequencingBacteroides, Veillonella, and FusobacteriumUpregulated
Escherichia-Shigella, Kluyvera, Fecalibacterium, Enterobacter, and DialisterDownregulated
Apostolou et al. (2011) [217]Reverse-transcription polymerase chain reactionStaphylococcus epidermidis, Streptococcus mitis, and Bacillus strainsUpregulated
Pancreatic Ductal Adenocarcinoma
Jesnowski et al. (2010) [165]Nested PCRHelicobacter pyloriNo expression
Ovarian Cancer
Chan et al. (1996) [218]Combined PCR-ELISA AssayMycoplasmaUpregulated
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Gupta, I.; Pedersen, S.; Vranic, S.; Al Moustafa, A.-E. Implications of Gut Microbiota in Epithelial–Mesenchymal Transition and Cancer Progression: A Concise Review. Cancers 2022, 14, 2964. https://doi.org/10.3390/cancers14122964

AMA Style

Gupta I, Pedersen S, Vranic S, Al Moustafa A-E. Implications of Gut Microbiota in Epithelial–Mesenchymal Transition and Cancer Progression: A Concise Review. Cancers. 2022; 14(12):2964. https://doi.org/10.3390/cancers14122964

Chicago/Turabian Style

Gupta, Ishita, Shona Pedersen, Semir Vranic, and Ala-Eddin Al Moustafa. 2022. "Implications of Gut Microbiota in Epithelial–Mesenchymal Transition and Cancer Progression: A Concise Review" Cancers 14, no. 12: 2964. https://doi.org/10.3390/cancers14122964

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

Article Metrics

Back to TopTop