Submitted Successfully!
To reward your contribution, here is a gift for you: A free trial for our video production service.
Thank you for your contribution! You can also upload a video entry or images related to this topic.
Version Summary Created by Modification Content Size Created at Operation
1 + 2965 word(s) 2965 2021-09-13 06:24:39 |
2 Format change -160 word(s) 2805 2021-11-15 02:47:00 | |
3 Format Change -25 word(s) 2780 2021-11-19 03:47:51 |

Video Upload Options

Do you have a full video?

Confirm

Are you sure to Delete?
Cite
If you have any further questions, please contact Encyclopedia Editorial Office.
Dupuy, B. Clostridioides Difficile Biofilm. Encyclopedia. Available online: https://encyclopedia.pub/entry/15944 (accessed on 27 July 2024).
Dupuy B. Clostridioides Difficile Biofilm. Encyclopedia. Available at: https://encyclopedia.pub/entry/15944. Accessed July 27, 2024.
Dupuy, Bruno. "Clostridioides Difficile Biofilm" Encyclopedia, https://encyclopedia.pub/entry/15944 (accessed July 27, 2024).
Dupuy, B. (2021, November 12). Clostridioides Difficile Biofilm. In Encyclopedia. https://encyclopedia.pub/entry/15944
Dupuy, Bruno. "Clostridioides Difficile Biofilm." Encyclopedia. Web. 12 November, 2021.
Clostridioides Difficile Biofilm
Edit

The microbiota inhabiting the intestinal tract provide several critical functions to its host. Microorganisms found at the mucosal layer form organized three-dimensional structures which are considered to be biofilms. Their development and functions are influenced by host factors, host-microbe interactions, and microbe-microbe interactions. These structures can dictate the health of their host by strengthening the natural defenses of the gut epithelium or cause disease by exacerbating underlying conditions. Biofilm communities can also block the establishment of pathogens and prevent infectious diseases. Although these biofilms are important for colonization resistance, new data provide evidence that gut biofilms can act as a reservoir for pathogens such as Clostridioides difficile.

dysbiosis mucosal-biofilm biofilm

1. Introduction

The human gastrointestinal tract (GIT) harbors a great diversity of microorganisms known as the gut microbiota [1][2]. The gut microbiota forms complex communities that coexist in an intimate relationship with the host, providing great benefits such as metabolic products and favoring the development of the immune system [3][4]. These gut microbial communities are present as planktonic cells or biofilm communities [5].

2. Health and Disease: Non-Invasive versus Invasive Gut Microbial Biofilms

In a healthy gut, a beneficial microbial biofilm formed by a complex ecological community will interact with the mucus layer and epithelium without invading the epithelia layer. This allows essential functions such as microbiota stability and resilience, which contribute to gut homeostasis and protect against infections [4][5]. Commensal biofilms offer a protective barrier against the proliferation and colonization of enteric pathogens, as well as of opportunistic pathobionts [6]. The resistance mechanisms offered by commensal communities against enteropathogens include the use of bacteriocins and short-chain fatty acids production, which inhibits the growth of pathogens and pathobionts [7][8][9]. Furthermore, commensal bacteria facilitate the host barrier function by thickening the mucus layer, inducing the expression of antimicrobial molecules and regulating the secretion of IgA [10][11][12][13]. Moreover, commensal microorganisms stimulate conversion of pro-IL-1β into active IL-1β [14] and induce the development of Th17 cells in the intestine, allowing protection against pathogens [15] (Figure 1).
Figure 1. Healthy microbiota biofilms versus a dysbiotic microbiota biofilms. In a healthy microbiota (left panel), the microbial density and diversity increase from the stomach to the colon. In the small intestine, biofilms are discontinuous and loose aggregates, while in the large intestine, biofilms are dense, continuous and attached to a uniform mucus layer (attached biofilms). The biofilms in the gut lumen are loosely attached to food particles or encapsulated in mucin (aggregate biofilms). Commensal biofilms facilitate the host barrier function by thickening the mucus layer, regulating the secretion of IgA, stimulating conversion of pro-IL-1β into active IL-1β and inducing the development of Th17 cells. A dysbiotic microbiota (right panel) presents (1) damaged mucus-biofilm exposing epithelium cells to luminal content or (2) invasive biofilms where bacteria come directly into contact with the epithelium. Both scenarios expose the intestinal epithelium to pathogens and pathobionts which can trigger an infection. Invasive polymicrobial biofilms could trigger cellular inflammation, abnormal cellular proliferation, increased epithelial permeability (activation of IL-6 and Stat3) in patients with colorectal cancer (CRC), increased IL-17 production and DNA damage in patients with familial adenomatous polyposis (FAP), and inflammatory bowel disease (IBD). Patients’ Adherent-invasive E. coli (AIEC) colonize the intestinal mucosa and stimulate the secretion of TNF-α and mucin degradation.
On the other hand, when dysbiosis occurs, the physiological conditions in the gut are altered, which affects the organisation of the mucosal biofilm. These changes can result in two possible scenarios: (1) the mucosal biofilm is damaged and forms aggregates of different sizes which leads to the exposure of epithelial cells to luminal content; or (2) an invasive biofilm is formed, bacteria colonize the inner sterile mucus layer and potentially come directly into contact with the epithelium (Figure 1). Both scenarios expose the intestinal epithelium to pathogens and pathobionts which can trigger an infection [5]. For example, changes in diet, such as fiber deficiency, promote the expansion of colonic mucus-degrading bacteria in mice, leading to the erosion of the colonic mucus barrier and facilitating the access to epithelial cells for enteric pathogens that cause colitis in mice such as Citrobacter rodentium [16], a surrogate pathogen for enterohemorrhagic E. coli (EHEC) and enteropathogenic E. coli (EPEC) [17].
Dysbiosis can also lead to invasive polymicrobial biofilms that induce cellular inflammation and abnormal cellular proliferation [18]. Invasive biofilms are associated directly with tumors. A signature of invasive biofilms is the reduction of E-cadherin on the surface of colonic epithelial cells and the high activation of IL-6 and Stat3, which increase epithelial permeability and tissue inflammation [18] (Figure 1). H. pylori is able to form biofilms in patients with peptic ulcer disease [19]. H. pylori forms biofilm-like microcolonies deep in the stomach glands and interacts directly with gastric progenitor and stem cells in tissues from mice and humans. These gland-associated bacteria accelerate stem cell proliferation and up-regulate the expression of stem cell–related genes, leading to glandular hyperplasia [20].
Bacterial biofilms present in the colon may also alter the host tissue microenvironment and induce metabolic changes in patients with colon cancer, as evident in metabolomic studies demonstrating changes in polyamine metabolite, including an upregulation of N1, N12-diacetylspermine. Increased polyamine concentrations are correlated with eukaryotic proliferation, potentially affecting cancer growth, development and progression [21].
Furthermore, invasive polymicrobial biofilms associated with diseases are composed of specific bacterial species or groups. For example, invasive biofilms associated with the colonic mucosa of familial adenomatous polyposis (FAP) patients, an inherited disorder characterized by cancer of the large intestine, were predominately composed of Escherichia coli and Bacteroides fragilis. These bacteria can secrete oncotoxins named colibactin (ClbB) and B. fragilis toxin (BFT), respectively, and these toxins were enriched in FAP patients. Furthermore, mice co-colonized with oncotoxin-producing strains had an increase in IL-17 production in the colon and increased DNA damage in colonic epithelial cells leading to faster onset of tumor [22] creation. Specifically, the BFT toxin triggers a pro-carcinogenic multi-step inflammatory cascade that increases the production of genotoxic oxygen radicals in colonic epithelial cells [23] (Figure 1).
Patients with colorectal cancer (CRC) have a higher number of Fusobacterium nucleatum and Streptococcus gallolyticus that surround the carcinoma or the adenoma tissues [24][25]. Both bacteria possess virulence factors that stimulate inflammatory and oncogenic responses [26]. Other bacteria that have been found in invasive biofilms in CRC patients are Campylobacter jejuni, Parvimonas micra, and Peptostreptococcus stomatis [27][28] (Figure 1).
Similarly, invasive biofilms are also associated with inflammatory bowel disease (IBD) such as Crohn’s disease (CD) and ulcerative colitis (UC) [29][30]. In patients with IBD, B. fragilis is responsible for more than 60% of the biofilm mass [31]. Another study found a high proportion of pro-inflammatory bacteria on the colonic mucosa of a young patient with ulcerative colitis such as Enterobacteriaceae, B. fragilis and P. aeruginosa [32]. Adherent-invasive E. coli (AIEC) were isolated from ileal biopsies of 36.4% of patients with CD. AIEC colonize the intestinal mucosa, survive and then replicate in epithelial cells and macrophages, which stimulate the secretion of large amounts of TNF-α [33] (Figure 1). Interestingly, AIEC possess a protease called Vat-AIEC that favors mucosa colonization by degrading mucins and decreasing mucus viscosity [34]. Also, an increased prevalence of mucolytic bacterial species such as Ruminococcus gnavus and Ruminococcus torques were found in CD and UC patients [35]. Furthermore, Enterococcus virulence factors were detected in children with IBD, and biofilm production was more frequent among Enterococcus strains isolated from children with IBD than in control strains [36] (Figure 1).
Overall, certain intestinal pathologies create an ideal environment which foster enrichment of specific bacterial groups. Bacteria associated with disease will form low diversity biofilm communities that exacerbate underlying conditions whereas bacteria associated with health will form a highly diverse biofilm community that strengthens the natural defenses of the gut epithelium [5]. Development and function of these biofilm communities will be influenced by host factors, host-microbe interactions, and microbe-microbe interactions [5].

3. Diversity of Interactions and Phenotypes in the Gut Biofilm Communities

Interactions in mixed-species biofilm communities of the gut can be neutral, positive or negative. Positive interactions are characterized by cooperation, commensalism and cross-feeding, whereas negative interactions are characterized by competition, exploitation and interference [37]. Cooperation involves one species that increases the fitness of another. Cooperation is not always reciprocal; however, if the interaction has a cost for one partner, an indirect benefit should be received for the interaction to be stable [38]. Competition is an indirect interaction between two species competing for a common resource. For example, Salmonella enterica induces inflammatory host responses that change the microbiota composition and suppress the microbiota’s growth [39]. In the case of exploitation, one species gains a fitness benefit at the cost to another, and this is also known as predation or parasitism [37]. Interference is a direct interaction where one species affects the fitness of another [37]. Interference includes the use of bacteriocins [40], type V, type VI and type VII secretion systems [41][42][43]. Overall, different types of interactions are occurring in biofilms and these will shape the properties and the special arrangement of biofilm communities.

4. Diverse Gut and Microbiota-Derived Signals Induce Biofilm Formation in Commensal Bacteria and Enteropathogens

The transition from a planktonic state to sessile growth is regulated by multiple steps and regulation cascades, and includes QS-dependent genes, the type IV pili (T4P), and the flagellum [44][45][46]. Biofilm formation is also guided by several environmental signals, which include mechanical signals, nutritional and metabolic cues, inorganic molecules, osmolarity, the presence of antimicrobial molecules, quorum-sensing derived signals, and host-derived signals [47].
Bacteria can initiate the transition from a planktonic state to biofilm in vivo to improve their survival against harmful conditions present in the host, to exploit a nutrient rich area that facilitates colonization, or to use the cooperative benefits of multicellular structures [48]. Biofilm formation can be controlled by stress response regulators that are activated by different stresses present in the host such as nutrient limitation, iron deprivation, sub-inhibitory concentrations of antibiotics, and osmotic stress [49][50][51]. Specific environmental conditions such as calcium concentration can increase the second messenger c-di-GMP concentrations that could trigger biofilm formation [52]. In some cases, biofilm formation is dependent on the nutritional conditions that will trigger metabolic adaptation and thus stimulate biofilm formation [46]. In the next section, we will focus on host-derived signals that induce biofilm formation in gut commensal bacteria and enteropathogens.

4.1. Host-Derived Factors and Biofilm Formation

Bile salts present in the intestinal tract of the host can induce biofilm formation in several enteropathogens and improve their survival against the toxic effects of bile [53]. Bile salts promote biofilm formation in V. cholerae by increasing the intracellular levels of c-di-GMP, which are caused by an increase in c-di-GMP synthesis by 3 diguanylate cyclases (DGCs) and decreased expression of one phosphodiesterase (PDE) [54]. The enteropathogen Shigella flexnerii also forms biofilm in response to the presence of deoxycholate (DOC), and this is mediated by the secreted protein IcsA, which is involved in cell-cell contacts and aggregative growth [55]. Similarly, vancomycin-resistant Enterococcus (VRE) is able to form biofilms in the presence of physiological concentrations of bile acids, which facilitates colonization and persistence. In VRE, the ability to form biofilms in response to bile salts is controlled by the histine kinase YycG/Walk of the WalRK two component system and the response regulator LiaR of the three-component regulatory system LiaFSR [56]. Likewise, B. fragilis treatment with bile salts increased bacterial co-aggregation, adhesion to intestinal epithelial cells and biofilm formation [57]. Exposure to bile salts induced morphological and transcriptional changes in B. fragilis, including overproduction of fimbria-like appendages and outer membrane vesicles, and increased expression of genes encoding RND-type efflux pumps and the major outer membrane protein, OmpA [57].
Additionally, Acinetobater baumannii, Cronobacter malonaticus, and Bifidobacterium formed more biofilms when exposed to bile salts [58][59][60]. In Bifidobacterium breve, bile-salt-induced biofilm formation involved QS, EPS production and eDNA release, and increased its viability when exposed to porcine bile salts [58]. In A. baumannii, exposure to bile salts increased expression of virulence factors associated with surface motility, biofilm, and type VI secretion systems, and these are also associated with activation of the QS system [59]. In the case of C. malonaticus, bile salts exposure induced an upregulation of the AcrAB-TolC system, but the molecular mechanisms involved in biofilm formation remain unknown [60].
When the commensal microbiota species B. breve and B. animalis were grown in taurocholic acid or porcine bile, the bacteria bound more effectively to mucin and formed more biofilm but the molecular mechanism is unknown [61]. Similarly, bile salts can induce biofilm formation in the commensal bacteria Bacteroides thetaiotaomicron, and this biofilm formation is dependent on the BT3563 DNAse that degrades extracellular DNA in the biofilm matrix [62].
Human secretory IgA (SlgA) appears to facilitate biofilm formation of the normal gut microbiota in vitro and of E. coli on the surface of cultured epithelial cells [63]. SlgA is a key factor that allows agglutination of bacteria and prevents their translocation to the gut epithelial cells, a process known as immune exclusion [64]. It was observed that mucin facilitated biofilm formation by E. coli by an unknown mechanism [63]. Similarly, type-2 mucin increased bacterial adhesion and biofilm formation in Listeria monocytogenes. This is mediated by the cell-surface protein InlL, which binds directly to Muc-2 [65]. Mucins are also used by C. jejuni as a signal to modulate the expression of virulence genes such as mucin degrading-enzymes, flagellin A and toxins [66]. Moreover, C jejuni is able to use fucose as a carbon source and shows chemotaxis towards fucose. C. jejuni biofilm formation decreased in the presence of fucose, suggesting that C. jejuni in a biofilm is able to coordinate fucose use based on its availability [67]. Mucus production in the colon is stimulated by the presence of hydrogen sulfide (H2S), which also promotes the establishment of biofilms in the GIT. H2S not only promoted biofilm formation by human microbiota ex vivo but also reduced the growth of planktonic bacteria [68].
Many studies have reported that several hormones and vitamins can affect biofilm formation and subsequent colonization. These factors include peptide hormones, steroid hormones such as catecholamine, and vitamin K [69]. For example, the hormone epinephrine was found to induce QS in EHEC [70]. In this study, a luxS deletion strain, which is unable to produce the EHEC autoinducer AI-3, responded to the host signal epinephrine and activated the expression of genes involved in biofilm formation [70]. Furthermore, E. coli biofilm formation is induced by insulin and is increased when glucose is present [71]. Indeed, the presence of insulin increased E. coli hydrophobicity and adherence to epithelial cells [72]. The gut commensal and opportunistic pathogen Enterococcus faecium can sense and respond to norepinephrine, a human hormone abundant in the gut, by inducing physiological changes, survival and colonization of the host tissues, and biofilm formation [73]. Catecholamines can also increase adhesion and biofilm formation in the enteropathogens Salmonella enteritidis and E. faecalis [74][75]. The specific molecular mechanisms of bacterial recognition of the hormones and the activation of regulatory pathways leading to increased biofilm formation have yet to be elucidated. Altogether, these studies show that there is cross-signaling between the host and the microbiota to allow maintenance of the gut homeostasis.

4.2. Antibiotics Affecting Biofilm Formation

Exposure to sub-inhibitory concentrations of antibiotics can induce or inhibit biofilm formation in bacteria. In E. faecalis, sub-inhibitory concentrations of tigecycline decrease biofilm formation [76], but sub-inhibitory concentrations of gentamicin significantly increased biofilm formation [77]. Similarly, sub-inhibitory concentrations of antibiotics that target the cell wall induced biofilm formation in E. faecalis [78]. The increase in biofilm formation was associated with an increase in cell lysis, extracellular DNA (eDNA) levels and cell density within the biofilm. This study included a mathematical model that predicted the changes in antibiotic-induced biofilms due to external alterations, showing that perturbations that reduce eDNA or decrease the number of living cells decreased biofilm induction, while compounds that increased cell lysis and cell wall inhibitors increased biofilm formation [78]. Similar results are also observed in gram-negative bacteria. For example, sub-inhibitory concentrations of aminoglycosides induced biofilm formation in E. coli [79]. However, sub-inhibitory concentrations of ceftazidime inhibited E. coli biofilm formation by increasing the extracellular concentration of indole [80].
Antibiotic resistance and tolerance can be mediated by efflux pumps and recent studies have suggested that efflux pumps may play a role in biofilm formation [81]. In E. coli, efflux pump genes such as isrA were highly expressed in biofilm bacteria compared to planktonic bacteria [82]. IsrA mediates the transport of the AI-2 signaling molecule involved in QS, suggesting that efflux pumps may play a role in the transport of the AI in E. coli biofilms, facilitating QS and promoting biofilm maturation [83]. Other multidrug efflux pumps such as AcrB and MdtABC were also involved in biofilm formation since corresponding mutant strains decreased biofilm formation and antibiotics resistance [84][85]. Similarly, the efflux pumps of S. enterica play an important role in biofilm formation. Indeed, the inactivation of efflux pumps inhibited the expression of the S. enterica curli, a surface protein filament that is an essential component of the biofilm matrix [86]. It was suggested that efflux pumps are involved in the activation of the regulator of curli gene expression [81]. In E. coli, some drug-induced stresses repressed production of curli and thus repressed biofilm formation [87].

References

  1. Eckburg, P.B. Diversity of the Human Intestinal Microbial Flora. Science 2005, 308, 1635–1638.
  2. Hooper, L.V.; Midtvedt, T.; Gordon, J.I. How Host-Microbial Interactions Shape the Nutrient Environment of the Mammalian Intestine. Annu. Rev. Nutr. 2002, 22, 283–307.
  3. Hooper, L.V.; Wong, M.H.; Thelin, A.; Hansson, L.; Falk, P.G.; Gordon, J.I. Molecular Analysis of Commensal Host-Microbial Relationships in the Intestine. Science 2001, 291, 881–884.
  4. Lozupone, C.A.; Stombaugh, J.I.; Gordon, J.I.; Jansson, J.K.; Knight, R. Diversity, Stability and Resilience of the Human Gut Microbiota. Nature 2012, 489, 220–230.
  5. Motta, J.-P.; Wallace, J.L.; Buret, A.G.; Deraison, C.; Vergnolle, N. Gastrointestinal Biofilms in Health and Disease. Nat. Rev. Gastroenterol. Hepatol. 2021, 18, 314–334.
  6. Kamada, N.; Chen, G.Y.; Inohara, N.; Núñez, G. Control of Pathogens and Pathobionts by the Gut Microbiota. Nat. Immunol. 2013, 14, 685–690.
  7. Fukuda, S.; Toh, H.; Hase, K.; Oshima, K.; Nakanishi, Y.; Yoshimura, K.; Tobe, T.; Clarke, J.M.; Topping, D.L.; Suzuki, T.; et al. Bifidobacteria Can Protect from Enteropathogenic Infection through Production of Acetate. Nature 2011, 469, 543–547.
  8. Hammami, R.; Fernandez, B.; Lacroix, C.; Fliss, I. Anti-Infective Properties of Bacteriocins: An Update. Cell. Mol. Life Sci. 2013, 70, 2947–2967.
  9. McDonald, J.A.K.; Mullish, B.H.; Pechlivanis, A.; Liu, Z.; Brignardello, J.; Kao, D.; Holmes, E.; Li, J.V.; Clarke, T.B.; Thursz, M.R.; et al. Inhibiting Growth of Clostridioides difficile by Restoring Valerate, Produced by the Intestinal Microbiota. Gastroenterology 2018, 155, 1495–1507.e15.
  10. Macpherson, A.J.; Geuking, M.B.; McCoy, K.D. Homeland Security: IgA Immunity at the Frontiers of the Body. Trends Immunol. 2012, 33, 160–167.
  11. Vaishnava, S.; Yamamoto, M.; Severson, K.M.; Ruhn, K.A.; Yu, X.; Koren, O.; Ley, R.; Wakeland, E.K.; Hooper, L.V. The Antibacterial Lectin RegIIIgamma Promotes the Spatial Segregation of Microbiota and Host in the Intestine. Science 2011, 334, 255–258.
  12. Vaishnava, S.; Behrendt, C.L.; Ismail, A.S.; Eckmann, L.; Hooper, L.V. Paneth Cells Directly Sense Gut Commensals and Maintain Homeostasis at the Intestinal Host-Microbial Interface. Proc. Natl. Acad. Sci. USA 2008, 105, 20858–20863.
  13. Wrzosek, L.; Miquel, S.; Noordine, M.-L.; Bouet, S.; Joncquel Chevalier-Curt, M.; Robert, V.; Philippe, C.; Bridonneau, C.; Cherbuy, C.; Robbe-Masselot, C.; et al. Bacteroides thetaiotaomicron and Faecalibacterium prausnitzii Influence the Production of Mucus Glycans and the Development of Goblet Cells in the Colonic Epithelium of a Gnotobiotic Model Rodent. BMC Biol. 2013, 11, 61.
  14. Franchi, L.; Kamada, N.; Nakamura, Y.; Burberry, A.; Kuffa, P.; Suzuki, S.; Shaw, M.H.; Kim, Y.-G.; Núñez, G. NLRC4-Driven Production of IL-1β Discriminates between Pathogenic and Commensal Bacteria and Promotes Host Intestinal Defense. Nat. Immunol. 2012, 13, 449–456.
  15. Ivanov, I.I.; Atarashi, K.; Manel, N.; Brodie, E.L.; Shima, T.; Karaoz, U.; Wei, D.; Goldfarb, K.C.; Santee, C.A.; Lynch, S.V.; et al. Induction of Intestinal Th17 Cells by Segmented Filamentous Bacteria. Cell 2009, 139, 485–498.
  16. Desai, M.S.; Seekatz, A.M.; Koropatkin, N.M.; Kamada, N.; Hickey, C.A.; Wolter, M.; Pudlo, N.A.; Kitamoto, S.; Terrapon, N.; Muller, A.; et al. A Dietary Fiber-Deprived Gut Microbiota Degrades the Colonic Mucus Barrier and Enhances Pathogen Susceptibility. Cell 2016, 167, 1339–1353.e21.
  17. Collins, J.W.; Keeney, K.M.; Crepin, V.F.; Rathinam, V.A.K.; Fitzgerald, K.A.; Finlay, B.B.; Frankel, G. Citrobacter rodentium: Infection, Inflammation and the Microbiota. Nat. Rev. Microbiol. 2014, 12, 612–623.
  18. Dejea, C.M.; Wick, E.C.; Hechenbleikner, E.M.; White, J.R.; Mark Welch, J.L.; Rossetti, B.J.; Peterson, S.N.; Snesrud, E.C.; Borisy, G.G.; Lazarev, M.; et al. Microbiota Organization Is a Distinct Feature of Proximal Colorectal Cancers. Proc. Natl. Acad Sci. USA 2014, 111, 18321–18326.
  19. Coticchia, J.M.; Sugawa, C.; Tran, V.R.; Gurrola, J.; Kowalski, E.; Carron, M.A. Presence and Density of Helicobacter pylori Biofilms in Human Gastric Mucosa in Patients with Peptic Ulcer Disease. J. Gastrointest. Surg. 2006, 10, 883–889.
  20. Sigal, M.; Rothenberg, M.E.; Logan, C.Y.; Lee, J.Y.; Honaker, R.W.; Cooper, R.L.; Passarelli, B.; Camorlinga, M.; Bouley, D.M.; Alvarez, G.; et al. Helicobacter pylori Activates and Expands Lgr5(+) Stem Cells Through Direct Colonization of the Gastric Glands. Gastroenterology 2015, 148, 1392–1404.e21.
  21. Johnson, C.H.; Dejea, C.M.; Edler, D.; Hoang, L.T.; Santidrian, A.F.; Felding, B.H.; Ivanisevic, J.; Cho, K.; Wick, E.C.; Hechenbleikner, E.M.; et al. Metabolism Links Bacterial Biofilms and Colon Carcinogenesis. Cell Metab. 2015, 21, 891–897.
  22. Dejea, C.M.; Fathi, P.; Craig, J.M.; Boleij, A.; Taddese, R.; Geis, A.L.; Wu, X.; DeStefano Shields, C.E.; Hechenbleikner, E.M.; Huso, D.L.; et al. Patients with Familial Adenomatous polyposis Harbor Colonic Biofilms Containing Tumorigenic Bacteria. Science 2018, 359, 592–597.
  23. Chung, L.; Thiele Orberg, E.; Geis, A.L.; Chan, J.L.; Fu, K.; DeStefano Shields, C.E.; Dejea, C.M.; Fathi, P.; Chen, J.; Finard, B.B.; et al. Bacteroides fragilis Toxin Coordinates a Pro-Carcinogenic Inflammatory Cascade via Targeting of Colonic Epithelial Cells. Cell Host Microbe 2018, 23, 203–214.e5.
  24. 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.
  25. Kumar, R.; Herold, J.L.; Schady, D.; Davis, J.; Kopetz, S.; Martinez-Moczygemba, M.; Murray, B.E.; Han, F.; Li, Y.; Callaway, E.; et al. Streptococcus gallolyticus Subsp. Gallolyticus Promotes Colorectal Tumor Development. PLoS Pathog. 2017, 13, e1006440.
  26. 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.
  27. Drewes, J.L.; White, J.R.; Dejea, C.M.; Fathi, P.; Iyadorai, T.; Vadivelu, J.; Roslani, A.C.; Wick, E.C.; Mongodin, E.F.; Loke, M.F.; et al. High-Resolution Bacterial 16S RRNA Gene Profile Meta-Analysis and Biofilm Status Reveal Common Colorectal Cancer Consortia. NPJ Biofilms Microbiomes 2017, 3, 34.
  28. 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.
  29. Macfarlane, S.; Furrie, E.; Cummings, J.H.; Macfarlane, G.T. Chemotaxonomic Analysis of Bacterial Populations Colonizing the Rectal Mucosa in Patients with Ulcerative Colitis. Clin. Infect. Dis. 2004, 38, 1690–1699.
  30. Swidsinski, A.; Loening-Baucke, V.; Herber, A. Mucosal Flora in Crohn’s Disease and Ulcerative Colitis—An Overview. J. Physiol. Pharmacol. 2009, 60 (Suppl. S6), 61–71.
  31. Swidsinski, A.; Weber, J.; Loening-Baucke, V.; Hale, L.P.; Lochs, H. Spatial Organization and Composition of the Mucosal Flora in Patients with Inflammatory Bowel Disease. J. Clin. Microbiol. 2005, 43, 3380–3389.
  32. Wang, M.; Molin, G.; Ahrné, S.; Adawi, D.; Jeppsson, B. High Proportions of Proinflammatory Bacteria on the Colonic Mucosa in a Young Patient with Ulcerative Colitis as Revealed by Cloning and Sequencing of 16S RRNA Genes. Dig. Dis. Sci. 2007, 52, 620–627.
  33. Rolhion, N.; Darfeuille-Michaud, A. Adherent-Invasive Escherichia coli in Inflammatory Bowel Disease. Inflamm. Bowel. Dis. 2007, 13, 1277–1283.
  34. Gibold, L.; Garenaux, E.; Dalmasso, G.; Gallucci, C.; Cia, D.; Mottet-Auselo, B.; Faïs, T.; Darfeuille-Michaud, A.; Nguyen, H.T.T.; Barnich, N.; et al. The Vat-AIEC Protease Promotes Crossing of the Intestinal Mucus Layer by Crohn’s Disease-Associated Escherichia coli. Cell Microbiol. 2016, 18, 617–631.
  35. Png, C.W.; Lindén, S.K.; Gilshenan, K.S.; Zoetendal, E.G.; McSweeney, C.S.; Sly, L.I.; McGuckin, M.A.; Florin, T.H.J. Mucolytic Bacteria with Increased Prevalence in IBD Mucosa Augment in vitro Utilization of Mucin by Other Bacteria. Am. J. Gastroenterol. 2010, 105, 2420–2428.
  36. Golińska, E.; Tomusiak, A.; Gosiewski, T.; Więcek, G.; Machul, A.; Mikołajczyk, D.; Bulanda, M.; Heczko, P.B.; Strus, M. Virulence Factors of Enterococcus Strains Isolated from Patients with Inflammatory Bowel Disease. World J. Gastroenterol. 2013, 19, 3562–3572.
  37. Coyte, K.Z.; Rakoff-Nahoum, S. Understanding Competition and Cooperation within the Mammalian Gut Microbiome. Curr. Biol. 2019, 29, R538–R544.
  38. Mitri, S.; Richard Foster, K. The Genotypic View of Social Interactions in Microbial Communities. Annu. Rev. Genet. 2013, 47, 247–273.
  39. Stecher, B.; Robbiani, R.; Walker, A.W.; Westendorf, A.M.; Barthel, M.; Kremer, M.; Chaffron, S.; Macpherson, A.J.; Buer, J.; Parkhill, J.; et al. Salmonella enterica Serovar Typhimurium Exploits Inflammation to Compete with the Intestinal Microbiota. PLoS Biol. 2007, 5, e244.
  40. Roelofs, K.G.; Coyne, M.J.; Gentyala, R.R.; Chatzidaki-Livanis, M.; Comstock, L.E. Bacteroidales Secreted Antimicrobial Proteins Target Surface Molecules Necessary for Gut Colonization and Mediate Competition In Vivo. mBio 2016, 7, e01055-16.
  41. Aoki, S.K.; Pamma, R.; Hernday, A.D.; Bickham, J.E.; Braaten, B.A.; Low, D.A. Contact-Dependent Inhibition of Growth in Escherichia coli. Science 2005, 309, 1245–1248.
  42. Sana, T.G.; Flaugnatti, N.; Lugo, K.A.; Lam, L.H.; Jacobson, A.; Baylot, V.; Durand, E.; Journet, L.; Cascales, E.; Monack, D.M. Salmonella typhimurium Utilizes a T6SS-Mediated Antibacterial Weapon to Establish in the Host Gut. Proc. Natl. Acad. Sci. USA 2016, 113, E5044–E5051.
  43. Whitney, J.C.; Peterson, S.B.; Kim, J.; Pazos, M.; Verster, A.J.; Radey, M.C.; Kulasekara, H.D.; Ching, M.Q.; Bullen, N.P.; Bryant, D.; et al. A Broadly Distributed Toxin Family Mediates Contact-Dependent Antagonism between Gram-Positive Bacteria. Elife 2017, 6, e26938.
  44. Barken, K.B.; Pamp, S.J.; Yang, L.; Gjermansen, M.; Bertrand, J.J.; Klausen, M.; Givskov, M.; Whitchurch, C.B.; Engel, J.N.; Tolker-Nielsen, T. Roles of Type IV Pili, Flagellum-Mediated Motility and Extracellular DNA in the Formation of Mature Multicellular Structures in Pseudomonas aeruginosa Biofilms. Environ. Microbiol. 2008, 10, 2331–2343.
  45. Inhülsen, S.; Aguilar, C.; Schmid, N.; Suppiger, A.; Riedel, K.; Eberl, L. Identification of Functions Linking Quorum Sensing with Biofilm Formation in Burkholderia cenocepacia H111. Microbiologyopen 2012, 1, 225–242.
  46. Tolker-Nielsen, T. Biofilm Development. Microbiol. Spectr. 2015, 3.
  47. Karatan, E.; Watnick, P. Signals, Regulatory Networks, and Materials That Build and Break Bacterial Biofilms. Microbiol. Mol. Biol. Rev. 2009, 73, 310–347.
  48. Jefferson, K.K. What Drives Bacteria to Produce a Biofilm? FEMS Microbiol. Lett. 2004, 236, 163–173.
  49. Gambino, M.; Cappitelli, F. Mini-Review: Biofilm Responses to Oxidative Stress. Biofouling 2016, 32, 167–178.
  50. Kang, D.; Kirienko, N.V. Interdependence between Iron Acquisition and Biofilm Formation in Pseudomonas aeruginosa. J. Microbiol. 2018, 56, 449–457.
  51. Kaplan, J.B. Antibiotic-Induced Biofilm Formation. Int. J. Artif. Organs. 2011, 34, 737–751.
  52. Chodur, D.M.; Coulter, P.; Isaacs, J.; Pu, M.; Fernandez, N.; Waters, C.M.; Rowe-Magnus, D.A. Environmental Calcium Initiates a Feed-Forward Signaling Circuit That Regulates Biofilm Formation and Rugosity in Vibrio Vulnificus. mBio 2018, 9, e01377-18.
  53. Hung, D.T.; Zhu, J.; Sturtevant, D.; Mekalanos, J.J. Bile Acids Stimulate Biofilm Formation in Vibrio Cholerae. Mol. Microbiol. 2006, 59, 193–201.
  54. Koestler, B.J.; Waters, C.M. Intestinal GPS: Bile and Bicarbonate Control Cyclic Di-GMP to Provide Vibrio Cholerae Spatial Cues within the Small Intestine. Gut Microbes 2014, 5, 775–780.
  55. Köseoğlu, V.K.; Hall, C.P.; Rodríguez-López, E.M.; Agaisse, H. The Autotransporter IcsA Promotes Shigella flexneri Biofilm Formation in the Presence of Bile Salts. Infect. Immun. 2019, 87, e00861-18.
  56. McKenney, P.T.; Yan, J.; Vaubourgeix, J.; Becattini, S.; Lampen, N.; Motzer, A.; Larson, P.J.; Dannaoui, D.; Fujisawa, S.; Xavier, J.B.; et al. Intestinal Bile Acids Induce a Morphotype Switch in Vancomycin-Resistant Enterococcus That Facilitates Intestinal Colonization. Cell Host Microbe 2019, 25, 695–705.e5.
  57. Pumbwe, L.; Skilbeck, C.A.; Nakano, V.; Avila-Campos, M.J.; Piazza, R.M.F.; Wexler, H.M. Bile Salts Enhance Bacterial Co-Aggregation, Bacterial-Intestinal Epithelial Cell Adhesion, Biofilm Formation and Antimicrobial Resistance of Bacteroides Fragilis. Microb. Pathog. 2007, 43, 78–87.
  58. Kelly, S.M.; Lanigan, N.; O’Neill, I.J.; Bottacini, F.; Lugli, G.A.; Viappiani, A.; Turroni, F.; Ventura, M.; van Sinderen, D. Bifidobacterial Biofilm Formation Is a Multifactorial Adaptive Phenomenon in Response to Bile Exposure. Sci. Rep. 2020, 10, 11598.
  59. López, M.; Blasco, L.; Gato, E.; Perez, A.; Fernández-Garcia, L.; Martínez-Martinez, L.; Fernández-Cuenca, F.; Rodríguez-Baño, J.; Pascual, A.; Bou, G.; et al. Response to Bile Salts in Clinical Strains of Acinetobacter baumannii Lacking the AdeABC Efflux Pump: Virulence Associated with Quorum Sensing. Front. Cell Infect. Microbiol. 2017, 7, 143.
  60. Wang, X.; Wang, Y.; Ling, N.; Shen, Y.; Zhang, D.; Liu, D.; Ou, D.; Wu, Q.; Ye, Y. Roles of TolC on Tolerance to Bile Salts and Biofilm Formation in Cronobacter malonaticus. J. Dairy Sci. 2021, 104, 9521–9531.
  61. Ambalam, P.; Kondepudi, K.K.; Nilsson, I.; Wadström, T.; Ljungh, A. Bile Enhances Cell Surface Hydrophobicity and Biofilm Formation of Bifidobacteria. Appl. Biochem. Biotechnol. 2014, 172, 1970–1981.
  62. Bechon, N.; Mihajlovic, J.; Lopes, A.-A.; Vendrell-Fernández, S.; Deschamps, J.; Briandet, R.; Sismeiro, O.; Martin-Verstraete, I.; Dupuy, B.; Ghigo, J.-M. Bacteroides thetaiotaomicron Uses a Widespread Extracellular DNase to Promote Bile-Dependent Biofilm Formation. bioRxiv 2021.
  63. Bollinger, R.R.; Everett, M.L.; Palestrant, D.; Love, S.D.; Lin, S.S.; Parker, W. Human Secretory Immunoglobulin A May Contribute to Biofilm Formation in the Gut. Immunology 2003, 109, 580–587.
  64. Everett, M.L.; Palestrant, D.; Miller, S.E.; Bollinger, R.R.; Parker, W. Immune Exclusion and Immune Inclusion: A New Model of Host-Bacterial Interactions in the Gut. Clin. Appl. Immunol. Rev. 2004, 4, 321–332.
  65. Popowska, M.; Krawczyk-Balska, A.; Ostrowski, R.; Desvaux, M. InlL from Listeria monocytogenes Is Involved in Biofilm Formation and Adhesion to Mucin. Front. Microbiol. 2017, 8, 660.
  66. Tu, Q.V.; McGuckin, M.A.; Mendz, G.L. Campylobacter jejuni Response to Human Mucin MUC2: Modulation of Colonization and Pathogenicity Determinants. J. Med. Microbiol. 2008, 57, 795–802.
  67. Dwivedi, R.; Nothaft, H.; Garber, J.; Xin Kin, L.; Stahl, M.; Flint, A.; van Vliet, A.H.M.; Stintzi, A.; Szymanski, C.M. L-Fucose Influences Chemotaxis and Biofilm Formation in Campylobacter jejuni. Mol. Microbiol. 2016, 101, 575–589.
  68. Motta, J.-P.; Flannigan, K.L.; Agbor, T.A.; Beatty, J.K.; Blackler, R.W.; Workentine, M.L.; Da Silva, G.J.; Wang, R.; Buret, A.G.; Wallace, J.L. Hydrogen Sulfide Protects from Colitis and Restores Intestinal Microbiota Biofilm and Mucus Production. Inflamm. Bowel. Dis. 2015, 21, 1006–1017.
  69. Feraco, D.; Blaha, M.; Khan, S.; Green, J.M.; Plotkin, B.J. Host Environmental Signals and Effects on Biofilm Formation. Microb. Pathog. 2016, 99, 253–263.
  70. Sperandio, V.; Torres, A.G.; Jarvis, B.; Nataro, J.P.; Kaper, J.B. Bacteria-Host Communication: The Language of Hormones. Proc. Natl. Acad. Sci. USA 2003, 100, 8951–8956.
  71. Plotkin, B.J.; Wu, Z.; Ward, K.; Nadella, S.; Green, J.M.; Rumnani, B. Effect of Human Insulin on the Formation of Catheter-Associated E. coli biofilms. OJU 2014, 4, 49–56.
  72. Klosowska, K.; Plotkin, B.J. Human Insulin Modulation of Escherichia coli Adherence and Chemotaxis. Am. J. Infect. Dis. 2006, 2, 197–200.
  73. Scardaci, R.; Varese, F.; Manfredi, M.; Marengo, E.; Mazzoli, R.; Pessione, E. Enterococcus faecium NCIMB10415 Responds to Norepinephrine by Altering Protein Profiles and Phenotypic Characters. J. Proteom. 2021, 231, 104003.
  74. Cambronel, M.; Nilly, F.; Mesguida, O.; Boukerb, A.M.; Racine, P.-J.; Baccouri, O.; Borrel, V.; Martel, J.; Fécamp, F.; Knowlton, R.; et al. Influence of Catecholamines (Epinephrine/Norepinephrine) on Biofilm Formation and Adhesion in Pathogenic and Probiotic Strains of Enterococcus faecalis. Front. Microbiol. 2020, 11, 1501.
  75. Hiller, C.C.; Lucca, V.; Carvalho, D.; Borsoi, A.; Borges, K.A.; Furian, T.Q.; do Nascimento, V.P. Influence of Catecholamines on Biofilm Formation by Salmonella enteritidis. Microb. Pathog. 2019, 130, 54–58.
  76. Maestre, J.R.; Aguilar, L.; Mateo, M.; Giménez, M.-J.; Méndez, M.-L.; Alou, L.; Granizo, J.-J.; Prieto, J. In vitro Interference of Tigecycline at Subinhibitory Concentrations on Biofilm Development by Enterococcus faecalis. J. Antimicrob. Chemother. 2012, 67, 1155–1158.
  77. Ozma, M.A.; Khodadadi, E.; Rezaee, M.A.; Kamounah, F.S.; Asgharzadeh, M.; Ganbarov, K.; Aghazadeh, M.; Yousefi, M.; Pirzadeh, T.; Kafil, H.S. Induction of Proteome Changes Involved in Biofilm Formation of Enterococcus faecalis in Response to Gentamicin. Microb. Pathog. 2021, 157, 105003.
  78. Yu, W.; Hallinen, K.M.; Wood, K.B. Interplay between Antibiotic Efficacy and Drug-Induced Lysis Underlies Enhanced Biofilm Formation at Subinhibitory Drug Concentrations. Antimicrob. Agents Chemother. 2018, 62, e01603-17.
  79. Hoffman, L.R.; D’Argenio, D.A.; MacCoss, M.J.; Zhang, Z.; Jones, R.A.; Miller, S.I. Aminoglycoside Antibiotics Induce Bacterial Biofilm Formation. Nature 2005, 436, 1171–1175.
  80. Sun, F.; Yuan, Q.; Wang, Y.; Cheng, L.; Li, X.; Feng, W.; Xia, P. Sub-Minimum Inhibitory Concentration Ceftazidime Inhibits Escherichia coli Biofilm Formation by Influencing the Levels of the IbpA Gene and Extracellular Indole. J. Chemother. 2020, 32, 7–14.
  81. Alav, I.; Sutton, J.M.; Rahman, K.M. Role of Bacterial Efflux Pumps in Biofilm Formation. J. Antimicrob. Chemother. 2018, 73, 2003–2020.
  82. Schembri, M.A.; Kjaergaard, K.; Klemm, P. Global Gene Expression in Escherichia coli Biofilms. Mol. Microbiol. 2003, 48, 253–267.
  83. Xavier, K.B.; Bassler, B.L. Regulation of Uptake and Processing of the Quorum-Sensing Autoinducer AI-2 in Escherichia coli. J. Bacteriol. 2005, 187, 238–248.
  84. Bay, D.C.; Stremick, C.A.; Slipski, C.J.; Turner, R.J. Secondary Multidrug Efflux Pump Mutants Alter Escherichia coli Biofilm Growth in the Presence of Cationic Antimicrobial Compounds. Res. Microbiol. 2017, 168, 208–221.
  85. Matsumura, K.; Furukawa, S.; Ogihara, H.; Morinaga, Y. Roles of Multidrug Efflux Pumps on the Biofilm Formation of Escherichia coli K-12. Biocontrol. Sci. 2011, 16, 69–72.
  86. Baugh, S.; Ekanayaka, A.S.; Piddock, L.J.V.; Webber, M.A. Loss of or Inhibition of All Multidrug Resistance Efflux Pumps of Salmonella enterica Serovar Typhimurium Results in Impaired Ability to Form a Biofilm. J. Antimicrob. Chemother. 2012, 67, 2409–2417.
  87. Teteneva, N.A.; Mart’yanov, S.V.; Esteban-López, M.; Kahnt, J.; Glatter, T.; Netrusov, A.I.; Plakunov, V.K.; Sourjik, V. Multiple Drug-Induced Stress Responses Inhibit Formation of Escherichia coli Biofilms. Appl. Environ. Microbiol. 2020, 86, e01113-20.
More
Information
Subjects: Biology
Contributor MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register :
View Times: 386
Revisions: 3 times (View History)
Update Date: 19 Nov 2021
1000/1000
Video Production Service