Skip to main content

SPECIALTY GRAND CHALLENGE article

Front. Cell. Infect. Microbiol., 22 February 2021
Sec. Biofilms
Volume 11 - 2021 | https://doi.org/10.3389/fcimb.2021.632429

Speciality Grand Challenge for “Biofilms”

  • 1Genetics of Biofilms Laboratory, Institut Pasteur, UMR CNRS2001, Paris, France
  • 2iîhree lnstitute, University of Technology Sydney, Sydney, NSW, Australia
  • 3Singapore Centre for Environmental Life Sciences Engineering, Nanyang Technological University, Singapore, Singapore


Introduction to Biofilms - Properties of Biofilms

The idea that most bacteria in nature are attached to surfaces was recognized as early as 1933 when Henrici stated after investigating fresh-water bacteria that “it is quite evident that for the most part water bacteria are not free floating organisms, but grow attached upon submerged surfaces” (Henrici, 1933). Zobell followed up with work that demonstrated seawater biofilms precede the attachment of other macro-fouling organisms (Zobell and Allen, 1935). Thus, began this field of research which was furthered by work on biofilms occurring in nature and the growing realization that a number of infections are biofilm-related (Costerton et al., 1999; Hall-Stoodley et al., 2004).

Biofilm Life Cycle

Biofilms are structurally diverse and dynamic (Hall-Stoodley et al., 2004). Biofilm formation occurs in stages beginning with the attachment of a cell on the surface, followed by division and secretion of extracellular polymeric substances (EPS) and formation of microcolonies (Costerton et al., 1987). There are channels that allow for exchange of nutrients and cellular waste between the microcolonies. The final stage of the biofilm lifecycle is dispersal which is driven by, but not limited to environmental cues and bacterially derived signals (reviewed in (McDougald et al., 2011). These signals include the second messenger cyclic-di-GMP (Jenal et al., 2017), nitric oxide (Barraud et al., 2009) and quorum sensing (Rice et al., 2005).

Increased Resistance to Antimicrobials and Matrix Properties

Biofilms are more resistant to a variety of stresses than their planktonic counterparts. For example, biofilms have been reported to be up to 1000-fold more resistant to antibiotics than planktonic cells (Costerton et al., 1999) as well as showing increased resistance to other environmental stresses and predation by protists (Matz et al., 2005). This increased resistance is due to changes in cellular metabolism of biofilm cells compared to planktonic cells (Lebeaux et al., 2014; Crabbé et al., 2019). It is also due to the physical protection provided by the extracellular matrix (Flemming and Wingender, 2010).

The biofilm matrix is composed of a variety of biopolymers, including polysaccharides, proteins, nucleic acids, lipids, and other biopolymers such as humic substances (Flemming and Wingender, 2010). These components can vary widely between different bacterial species. The polysaccharides and proteins help to protect biofilm cells from antimicrobials by acting as a barrier to their penetration into the biofilm. Pseudomonas aeruginosa requires extracellular DNA or eDNA to form a biofilm and for motility within the biofilm (Whitchurch et al., 2002; Gloag et al., 2013). It is this matrix that lends stability to the biofilm making them difficult to remove. The matrix can also bind extracellular enzymes making them available to biofilm cells and sequester other damaging compounds.

Challenges in Methods for Investigating Biofilms

From Monospecies to Polymicrobial Biofilms

Since the early observation of biofilms in their natural environments, notably in rivers, researchers have realized the heterogeneity and complexity of biofilm composition and structure. After a long period during which monospecies biofilms have been studied with success we have moved towards the study of polymicrobial biofilms, which are also relevant in multiple biofilm-associated infections in animals and humans and often at the origin of chronic infections (Peters et al., 2012; Wolcott et al., 2013). Among the multitude of in vitro and in vivo models of biofilm formation and infection models that have been developed (Coenye and Nelis, 2010; Lebeaux et al., 2013; Brackman and Coenye, 2016) some specific models are dedicated to polymicrobial biofilms (Gabrilska and Rumbaugh, 2015; Tay et al., 2016). The relevance of studying polymicrobial biofilms was demonstrated by the importance of the interactions of micro-organisms for the development of biofilms in the context of infection, notably through the cross-talk between species and their impact on mechanisms of antimicrobial recalcitrance (Kragh et al., 2016; Liu et al., 2016; Orazi and O’Toole, 2019; Ibberson and Whiteley, 2020). Further developments of such models are essential to uncover the molecular mechanisms beyond the specific behavior of micro-organisms during such polymicrobial biofilms.

Tackling the Heterogeneity of Biofilms

In addition to the complexity of polymicrobial biofilms, one hallmark of biofilms in general is their heterogeneity in terms of local environmental conditions that translates into a heterogeneity of physiology of biofilm cells. While studies using technologies such as transcriptomics, proteomics and metabolomics have been successfully applied to image the global physiology or responses to stress of biofilms cells (Seneviratne et al., 2020), there is a clear need to think singularity and to develop technologies that would allow for the study of the physiology of individual biofilm cells and to understand the social interactions within biofilms (Rode et al., 2020). These studies can be performed with the improvement of methods such as single-cell RNA transcriptomics (Blattman et al., 2020; Imdahl et al., 2020) and technologies used to isolate cells from biofilms (Ma et al., 2019). Such recently developed technologies have been successfully applied to image the physiology of S. aureus persister cells within macrophages and should be applied hopefully to biofilms. Imaging technologies have also recently improved to locate and trace individual cells within biofilms. Multiplex FISH technology has been applied to characterize the biogeography of the oral microbiota (Valm et al., 2012; Valm, 2019) while recent microscopy imaging and analyzing technologies allowing for dynamically tracking single biofilm bacteria (Hartmann et al., 2019), for example upon antibiotic stress in V. cholerae (Díaz-Pascual et al., 2019). Additionally, some very promising tools have been recently adapted to metabolite identification at the single cell level in biofilm, for instance using Raman spectroscopy, Mass spectrometry or electro-chemical chip/fluorophores (Baig et al., 2016; Bellin et al., 2016; Bodelón et al., 2016; Schiessl et al., 2019; Geier et al., 2020; Yang et al., 2020). The upcoming challenge will be, in addition to further development of these technologies, to adapt them to direct observation or characterization of in situ biofilms and to models, whether in vitro, ex vivo, or in vivo, that better feature real in situ biofilms.

Going Toward a Better Understanding of In Vivo Biofilms: Improving Detection and Observation of In Situ Biofilms and Developing Novel Models?

Today there is absolutely no doubt that biofilms occur in vivo, especially in compromised hosts, with almost 80% of human infections now recognized to be biofilm-related (NIH, 2002). Biofilms form in body compartments as well as attached to cells and indwelling devices (Bjarnsholt et al., 2013). One well-studied example of this is the infection of cystic fibrosis (CF) patients’ lungs by P. aeruginosa. These biofilms are extremely resistant to antimicrobials for reasons discussed above but also due to their ability to resist host defenses. For example, in both in vitro and in vivo CF models, P. aeruginosa was shown to respond to the presence of host immune cells by upregulating a number of quorum-sensing regulated virulence factors such as rhamnolipids which are known to kill immune cells (Alhede et al., 2009). Other common chronic infections include chronic otitis media and chronic wounds that tend to be caused by multispecies infections (Burmølle et al., 2010). Other major sources of in vivo biofilms are indwelling medical devices such as intravenous or urinary catheters, pacemakers, endotracheal tubes and artificial joints (Lebeaux et al., 2014; Stewart and Bjarnsholt, 2020). Microorganisms commonly associated with medical devices are Staphylococcus epidermidis, Staphylococcus aureus, and P. aeruginosa (Hall-Stoodley et al., 2004). These biofilms often result in replacement of the device, which for artificial joints is a significant task. The dispersal of these biofilms has been discussed as a potential treatment, however a large number of planktonic cells entering the blood stream may be more dangerous than other treatments. Numerous infections are also related to biofilms in animals, for example in bovine mastitis, pneumonia, liver abscess, enteritis, urinary tract infection, otitis, wound infection, etc. (Abdullahi et al., 2016) and in plants where bacterial and fungal pathogens colonization can be at the origin of different diseases (Bogino et al., 2013; Castiblanco and Sundin, 2016; Motaung et al., 2020).

For decades our vision of how biofilms look has been governed by our knowledge developed from the use of in vitro biofilms models. As stated above, the increasing number of biofilms isolated from real infection situations have brought into question this conventional thinking with the recognition that, with some exception, in vivo biofilms rather correspond to patches or cell aggregates that rarely exceed 200 µm (Bjarnsholt et al., 2013) with important consequences on how biofilms finally interact with the host (Alhede et al., 2020). These observations have led to several important statements: in most cases, in vitro, and sometimes in vivo biofilms models do not relevantly recapitulate biofilms in their real infectious context; the diagnosis of the presence of biofilms linked to infection cannot be made a posteriori if we want clinicians to adapt their treatment to the presence of a biofilm. With these statements come potential actions: on one side the necessity to improve detection methods that could, for some of them, allow early detection of biofilms in the host and, for others, give a better characterization of the environment of infectious biofilms. Thus, the need to implement some of the in situ information into the current in vitro and in vivo models, and to develop novel models is paramount. It is clear that to this day some effort has been made in these directions, but we will have to do much more to reach these objectives.

We are still trying to identify biofilm biomarkers that could be used not only for detection but also for vaccine development. Important work has been done in the laboratory of the late Prof. Mark Shirliff in this direction notably with a recent demonstration that S. aureus orthopedic implant infections could be detected in human sinuvial fluids using an antibody against the manganese transporter MntC (Harro et al., 2019; Harro et al., 2020). Peptide-fluorescently labeled probes have been recently identified and used to specifically label P. aeruginosa biofilms in vivo (Locke et al., 2020). Additionally, numerous other detection methods are under development and efforts should be made to allow these technologies to translate into the clinic (Achinas et al., 2020; Parlak and Richter-Dahlfors, 2020). Among them, recently there has been the successful use of imaging technologies using bioluminescence (Chauhan et al., 2016; Hoffmann et al., 2019; Maiden et al., 2019; Gordon et al., 2020; Kreth et al., 2020; Redman et al., 2020; Van Dyck et al., 2020) or intravital imaging (Thanabalasuriar et al., 2019; Abdul Hamid et al., 2020; Gries et al., 2020; Tian et al., 2020) to study the effect of antimicrobial agents on biofilms or the behavior of immune cells in contact with biofilms in various in vivo models of biofilm-related infections.

The concept of the micro-environment of biofilm cells during infection has been well conceptualized in a recent perspective article on chronic wounds where several defined zones can be identified and on which information should be gathered in order to access to a global view of the biofilm infectious micro-environment (Kirketerp-Møller et al., 2020). Ideally, non-destructive methods to sample and analyze this micro-environment using micro-dialysis, micro-probes or electrochemical probes should be adapted to in vivo measurement as recently explained in (Røder et al., 2020), however, some of these methods also have applications for biofilm detection.

Improving the models for a better understanding of biofilm formation and biofilm cells behavior is also a major objective that should be reached by researchers in the field. Initiatives to evaluate the available models are welcome and should provide a better view of the current more relevant models with some clues on how to improve them (Cornforth et al., 2020). With the increased pressure to restrict the use of animals in research there is also a dire need to identify surrogate models. While there is some utility to pursue the development of non-mammalian models of biofilm-infection such as Drosophila melanogaster, Galleria mellonella or Danio rerio (Zebrafish) (Lebeaux et al., 2013), identification of ex vivo models that recreate the infectious environment should be encouraged. Among recently developed ex vivo models are, for example, an ex vivo pig lung biofilm model used for understanding antibiotic tolerance of P. aeruginosa biofilms (Hassan et al., 2020) or an ex vivo murine skin biopsy model to characterize B. burgdoferi biofilms, the etiological agent of Lyme disease (Torres et al., 2020). Last but not least with the development of reconstituted organs such as organoids or organ-on-a-chip, one could anticipate that adaptations of these models to the understanding of the physiopathology of biofilm infections, interactions of biofilms with the immune response or biofilm behaviors in the presence of antimicrobials would provide important information relevant to biofilms in infectious contexts (Jimi et al., 2017; Choi et al., 2020; Yuan et al., 2020).

Major Challenges

As described above, the field of biofilms and interactions with higher organisms has clear challenges for future research. 1) The development of new and improved multi-species biofilm models will allow us to investigate interactions that occur in these biofilms, as most in vivo biofilms are multispecies biofilms. This is clearly the next step in understanding biofilms in nature and in disease contexts, and this should come with the use of technologies that would allow probing the infectious micro-environment of biofilms. 2) The development of more relevant models, whether in vitro, ex vivo, or in vivo, to elucidate interactions of biofilms with the host in more detail. These models will be necessary not only for understanding biofilm host interactions but also for the investigation of eradication methods or compounds in a more real setting. 3) The development of imaging and detection technologies that could be used to better visualize and detect biofilms in situ and should allow when necessary to access the physiology on individual cells within biofilms. As new methods are developed, using these for biofilm research will generate new knowledge and lead to new areas of biofilm research.

Author Contributions

CB and DM wrote the manuscript. All authors contributed to the article and approved the submitted version.

Conflict of Interest

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

Abdul Hamid A. I., Nakusi L., Givskov M., Chang Y. T., Marquès C., Gueirard P. (2020). A mouse ear skin model to study the dynamics of innate immune responses against Staphylococcus aureus biofilms. BMC Microbiol. 20, 22. doi: 10.1186/s12866-019-1635-z

PubMed Abstract | CrossRef Full Text | Google Scholar

Abdullahi U. F., Igwenagu E., Mu’azu A., Aliyu S., Umar M. I. (2016). Intrigues of biofilm: A perspective in veterinary medicine. Vet. World 9, 12–18. doi: 10.14202/vetworld.2016.12-18

PubMed Abstract | CrossRef Full Text | Google Scholar

Achinas S., Yska S. K., Charalampogiannis N., Krooneman J., Euverink G. J. W. (2020). A Technological Understanding of Biofilm Detection Techniques: A Review. Mater. (Basel) 13(14):3417. doi: 10.3390/ma13143147

CrossRef Full Text | Google Scholar

Alhede M., Bjarnsholt T., Jensen P., Phipps R. K., Moser C., Christophersen L., et al. (2009). Pseudomonas aeruginosa recognizes and responds aggressively to the presence of polymorphonuclear leukocytes. Microbiol. (Reading) 155, 3500–3508. doi: 10.1099/mic.0.031443-0

CrossRef Full Text | Google Scholar

Alhede M., Alhede M., Qvortrup K., Kragh K. N., Jensen P., Stewart P. S., et al. (2020). The origin of extracellular DNA in bacterial biofilm infections in vivo. Pathog. Dis. 78(2):ftaa018. doi: 10.1093/femspd/ftaa018

PubMed Abstract | CrossRef Full Text | Google Scholar

Baig N., Polisetti S., Morales-Soto N., Dunham S. J. B., Sweedler J. V., Shrout J. D., et al. (2016). Label-free molecular imaging of bacterial communities of the opportunistic pathogen Pseudomonas aeruginosa. Proc. SPIE Int. Soc. Opt. Eng. 9930:993004. doi: 10.1117/12.2236695

PubMed Abstract | CrossRef Full Text | Google Scholar

Barraud N., Storey M. V., Moore Z. P., Webb J. S., Rice S. A., Kjelleberg S. (2009). Nitric oxide-mediated dispersal in single- and multi-species biofilms of clinically and industrially relevant microorganisms. Microb. Biotechnol. 2, 370–378. doi: 10.1111/j.1751-7915.2009.00098.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Bellin D. L., Sakhtah H., Zhang Y., Price-Whelan A., Dietrich L. E., Shepard K. L. (2016). Electrochemical camera chip for simultaneous imaging of multiple metabolites in biofilms. Nat. Commun. 7, 10535. doi: 10.1038/ncomms10535

PubMed Abstract | CrossRef Full Text | Google Scholar

Bjarnsholt T., Alhede M., Alhede M., Eickhardt-Sørensen S. R., Moser C., Kühl M., et al. (2013). The in vivo biofilm. Trends Microbiol. 21, 466–474. doi: 10.1016/j.tim.2013.06.002

PubMed Abstract | CrossRef Full Text | Google Scholar

Blattman S. B., Jiang W., Oikonomou P., Tavazoie S. (2020). Prokaryotic single-cell RNA sequencing by in situ combinatorial indexing. Nat. Microbiol. 5, 1192–1201. doi: 10.1038/s41564-020-0729-6

PubMed Abstract | CrossRef Full Text | Google Scholar

Bodelón G., Montes-García V., López-Puente V., Hill E. H., Hamon C., Sanz-Ortiz M. N., et al. (2016). Detection and imaging of quorum sensing in Pseudomonas aeruginosa biofilm communities by surface-enhanced resonance Raman scattering. Nat. Mater. 15, 1203–1211. doi: 10.1038/nmat4720

PubMed Abstract | CrossRef Full Text | Google Scholar

Bogino P. C., Oliva Mde L., Sorroche F. G., Giordano W. (2013). The role of bacterial biofilms and surface components in plant-bacterial associations. Int. J. Mol. Sci. 14, 15838–15859. doi: 10.3390/ijms140815838

PubMed Abstract | CrossRef Full Text | Google Scholar

Brackman G., Coenye T. (2016). In Vitro and In Vivo Biofilm Wound Models and Their Application. Adv. Exp. Med. Biol. 897, 15–32. doi: 10.1007/5584_2015_5002

PubMed Abstract | CrossRef Full Text | Google Scholar

Burmølle M., Thomsen T. R., Fazli M., Dige I., Christensen L., Homøe P., et al. (2010). Biofilms in chronic infections - a matter of opportunity - monospecies biofilms in multispecies infections. FEMS Immunol. Med. Microbiol. 59, 324–336. doi: 10.1111/j.1574-695X.2010.00714.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Castiblanco L. F., Sundin G. W. (2016). New insights on molecular regulation of biofilm formation in plant-associated bacteria. J. Integr. Plant Biol. 58, 362–372. doi: 10.1111/jipb.12428

PubMed Abstract | CrossRef Full Text | Google Scholar

Chauhan A., Ghigo J. M., Beloin C. (2016). Study of in vivo catheter biofilm infections using pediatric central venous catheter implanted in rat. Nat. Protoc. 11, 525–541. doi: 10.1038/nprot.2016.033

PubMed Abstract | CrossRef Full Text | Google Scholar

Choi K. G., Wu B. C., Lee A. H., Baquir B., Hancock R. E. W. (2020). Utilizing Organoid and Air-Liquid Interface Models as a Screening Method in the Development of New Host Defense Peptides. Front. Cell Infect. Microbiol. 10, 228. doi: 10.3389/fcimb.2020.00228

PubMed Abstract | CrossRef Full Text | Google Scholar

Coenye T., Nelis H. J. (2010). In vitro and in vivo model systems to study microbial biofilm formation. J. Microbiol. Methods 83, 89–105. doi: 10.1016/j.mimet.2010.08.018

PubMed Abstract | CrossRef Full Text | Google Scholar

Cornforth D. M., Diggle F. L., Melvin J. A., Bomberger J. M., Whiteley M. (2020). Quantitative Framework for Model Evaluation in Microbiology Research Using Pseudomonas aeruginosa and Cystic Fibrosis Infection as a Test Case. mBio 11(1):e03042-19. doi: 10.1128/mBio.03042-19

PubMed Abstract | CrossRef Full Text | Google Scholar

Costerton J. W., Cheng K. J., Geesey G. G., Ladd T. I., Nickel J. C., Dasgupta M., et al. (1987). Bacterial biofilms in nature and disease. Annu. Rev. Microbiol. 41, 435–464. doi: 10.1146/annurev.mi.41.100187.002251

PubMed Abstract | CrossRef Full Text | Google Scholar

Costerton J. W., Stewart P. S., Greenberg E. P. (1999). Bacterial biofilms: a common cause of persistent infections. Science 284, 1318–1322. doi: 10.1126/science.284.5418.1318

PubMed Abstract | CrossRef Full Text | Google Scholar

Crabbé A., Jensen P., Bjarnsholt T., Coenye T. (2019). Antimicrobial Tolerance and Metabolic Adaptations in Microbial Biofilms. Trends Microbiol. 27, 850–863. doi: 10.1016/j.tim.2019.05.003

PubMed Abstract | CrossRef Full Text | Google Scholar

Díaz-Pascual F., Hartmann R., Lempp M., Vidakovic L., Song B., Jeckel H., et al. (2019). Breakdown of Vibrio cholerae biofilm architecture induced by antibiotics disrupts community barrier function. Nat. Microbiol. 4, 2136–2145. doi: 10.1038/s41564-019-0579-2

PubMed Abstract | CrossRef Full Text | Google Scholar

Flemming H. C., Wingender J. (2010). The biofilm matrix. Nat. Rev. Microbiol. 8, 623–633. doi: 10.1038/nrmicro2415

PubMed Abstract | CrossRef Full Text | Google Scholar

Gabrilska R. A., Rumbaugh K. P. (2015). Biofilm models of polymicrobial infection. Future Microbiol. 10, 1997–2015. doi: 10.2217/fmb.15.109

PubMed Abstract | CrossRef Full Text | Google Scholar

Geier B., Sogin E. M., Michellod D., Janda M., Kompauer M., Spengler B., et al. (2020). Spatial metabolomics of in situ host-microbe interactions at the micrometre scale. Nat. Microbiol. 5, 498–510. doi: 10.1038/s41564-019-0664-6

PubMed Abstract | CrossRef Full Text | Google Scholar

Gloag E. S., Turnbull L., Huang A., Vallotton P., Wang H., Nolan L. M., et al. (2013). Self-organization of bacterial biofilms is facilitated by extracellular DNA. Proc. Natl. Acad. Sci. U.S.A. 110, 11541–11546. doi: 10.1073/pnas.1218898110

PubMed Abstract | CrossRef Full Text | Google Scholar

Gordon O., Miller R. J., Thompson J. M., Ordonez A. A., Klunk M. H., Dikeman D. A., et al. (2020). Rabbit model of Staphylococcus aureus implant-associated spinal infection. Dis. Model Mech. 13(7):dmm045385. doi: 10.1242/dmm.045385

PubMed Abstract | CrossRef Full Text | Google Scholar

Gries C. M., Rivas Z., Chen J., Lo D. D. (2020). Intravital Multiphoton Examination of Implant-Associated Staphylococcus aureus Biofilm Infection. Front. Cell Infect. Microbiol. 10, 574092. doi: 10.3389/fcimb.2020.574092

PubMed Abstract | CrossRef Full Text | Google Scholar

Hall-Stoodley L., Costerton J. W., Stoodley P. (2004). Bacterial biofilms: from the natural environment to infectious diseases. Nat. Rev. Microbiol. 2, 95–108. doi: 10.1038/nrmicro821

PubMed Abstract | CrossRef Full Text | Google Scholar

Harro J. M., Achermann Y., Freiberg J. A., Allison D. L., Brao K. J., Marinos D. P., et al. (2019). Clearance of Staphylococcus aureus from In Vivo Models of Chronic Infection by Immunization Requires Both Planktonic and Biofilm Antigens. Infect. Immun. 88(1):e00586-19. doi: 10.1128/IAI.00586-19

PubMed Abstract | CrossRef Full Text | Google Scholar

Harro J. M., Shirtliff M. E., Arnold W., Kofonow J. M., Dammling C., Achermann Y., et al. (2020). Development of a Novel and Rapid Antibody-Based Diagnostic for Chronic Staphylococcus aureus Infections Based on Biofilm Antigens. J. Clin. Microbiol. 58(5):e01414-19. doi: 10.1128/JCM.01414-19

PubMed Abstract | CrossRef Full Text | Google Scholar

Hartmann R., Singh P. K., Pearce P., Mok R., Song B., Díaz-Pascual F., et al. (2019). Emergence of three-dimensional order and structure in growing biofilms. Nat. Phys. 15, 251–256. doi: 10.1038/s41567-018-0356-9

PubMed Abstract | CrossRef Full Text | Google Scholar

Hassan M. M., Harrington N. E., Sweeney E., Harrison F. (2020). Predicting Antibiotic-Associated Virulence of Pseudomonas aeruginosa Using an ex vivo Lung Biofilm Model. Front. Microbiol. 11, 568510. doi: 10.3389/fmicb.2020.568510

PubMed Abstract | CrossRef Full Text | Google Scholar

Henrici A. T. (1933). Studies of Freshwater Bacteria: I. A Direct Microscopic Technique. J. Bacteriol. 25, 277–287. doi: 10.1128/JB.25.3.277-287.1933

PubMed Abstract | CrossRef Full Text | Google Scholar

Hoffmann J. P., Friedman J. K., Wang Y., Mclachlan J. B., Sammarco M. C., Morici L. A., et al. (2019). In situ Treatment With Novel Microbiocide Inhibits Methicillin Resistant Staphylococcus aureus in a Murine Wound Infection Model. Front. Microbiol. 10, 3106. doi: 10.3389/fmicb.2019.03106

PubMed Abstract | CrossRef Full Text | Google Scholar

Ibberson C. B., Whiteley M. (2020). The social life of microbes in chronic infection. Curr. Opin. Microbiol. 53, 44–50. doi: 10.1016/j.mib.2020.02.003

PubMed Abstract | CrossRef Full Text | Google Scholar

Imdahl F., Vafadarnejad E., Homberger C., Saliba A. E., Vogel J. (2020). Single-cell RNA-sequencing reports growth-condition-specific global transcriptomes of individual bacteria. Nat. Microbiol. 5, 1202–1206. doi: 10.1038/s41564-020-0774-1

PubMed Abstract | CrossRef Full Text | Google Scholar

Jenal U., Reinders A., Lori C. (2017). Cyclic di-GMP: second messenger extraordinaire. Nat. Rev. Microbiol. 15, 271–284. doi: 10.1038/nrmicro.2016.190

PubMed Abstract | CrossRef Full Text | Google Scholar

Jimi S., Miyazaki M., Takata T., Ohjimi H., Akita S., Hara S. (2017). Increased drug resistance of meticillin-resistant Staphylococcus aureus biofilms formed on a mouse dermal chip model. J. Med. Microbiol. 66, 542–550. doi: 10.1099/jmm.0.000461

PubMed Abstract | CrossRef Full Text | Google Scholar

Kirketerp-Møller K., Stewart P. S., Bjarnsholt T. (2020). The zone model: A conceptual model for understanding the microenvironment of chronic wound infection. Wound Repair Regener. 28, 593–599. doi: 10.1111/wrr.12841

CrossRef Full Text | Google Scholar

Kragh K. N., Hutchison J. B., Melaugh G., Rodesney C., Roberts A. E., Irie Y., et al. (2016). Role of Multicellular Aggregates in Biofilm Formation. mBio 7, e00237. doi: 10.1128/mBio.00237-16

PubMed Abstract | CrossRef Full Text | Google Scholar

Kreth J., Abdelrahman Y. M., Merritt J. (2020). Multiplex Imaging of Polymicrobial Communities-Murine Models to Study Oral Microbiome Interactions. Methods Mol. Biol. 2081, 107–126. doi: 10.1007/978-1-4939-9940-8_8

PubMed Abstract | CrossRef Full Text | Google Scholar

Lebeaux D., Chauhan A., Rendueles O., Beloin C. (2013). From in vitro to in vivo Models of Bacterial Biofilm-Related Infections. Pathogens 2, 288–356. doi: 10.3390/pathogens2020288

PubMed Abstract | CrossRef Full Text | Google Scholar

Lebeaux D., Ghigo J. M., Beloin C. (2014). Biofilm-related infections: bridging the gap between clinical management and fundamental aspects of recalcitrance toward antibiotics. Microbiol. Mol. Biol. Rev. 78, 510–543. doi: 10.1128/MMBR.00013-14

PubMed Abstract | CrossRef Full Text | Google Scholar

Liu W., Røder H. L., Madsen J. S., Bjarnsholt T., Sørensen S. J., Burmølle M. (2016). Interspecific Bacterial Interactions are Reflected in Multispecies Biofilm Spatial Organization. Front. Microbiol. 7, 1366. doi: 10.3389/fmicb.2016.01366

PubMed Abstract | CrossRef Full Text | Google Scholar

Locke L. W., Shankaran K., Gong L., Stoodley P., Vozar S. L., Cole S. L., et al. (2020). Evaluation of Peptide-Based Probes toward In Vivo Diagnostic Imaging of Bacterial Biofilm-Associated Infections. ACS Infect. Dis. 6, 2086–2098. doi: 10.1021/acsinfecdis.0c00125

PubMed Abstract | CrossRef Full Text | Google Scholar

Ma Q., Bücking H., Gonzalez Hernandez J. L., Subramanian S. (2019). Single-Cell RNA Sequencing of Plant-Associated Bacterial Communities. Front. Microbiol. 10, 2452. doi: 10.3389/fmicb.2019.02452

PubMed Abstract | CrossRef Full Text | Google Scholar

Maiden M. M., Zachos M. P., Waters C. M. (2019). Hydrogels Embedded With Melittin and Tobramycin Are Effective Against Pseudomonas aeruginosa Biofilms in an Animal Wound Model. Front. Microbiol. 10, 1348. doi: 10.3389/fmicb.2019.01348

PubMed Abstract | CrossRef Full Text | Google Scholar

Matz C., Mcdougald D., Moreno A. M., Yung P. Y., Yildiz F. H., Kjelleberg S. (2005). Biofilm formation and phenotypic variation enhance predation-driven persistence of Vibrio cholerae. Proc. Natl. Acad. Sci. U.S.A. 102, 16819–16824. doi: 10.1073/pnas.0505350102

PubMed Abstract | CrossRef Full Text | Google Scholar

McDougald D., Rice S. A., Barraud N., Steinberg P. D., Kjelleberg S. (2011). Should we stay or should we go: mechanisms and ecological consequences for biofilm dispersal. Nat. Rev. Microbiol. 10, 39–50. doi: 10.1038/nrmicro2695

PubMed Abstract | CrossRef Full Text | Google Scholar

Motaung T. E., Peremore C., Wingfield B., Steenkamp E. (2020). Plant-associated fungal biofilms - knowns and unknowns. FEMS Microbiol. Ecol. 96(12):fiaa224. doi: 10.1093/femsec/fiaa224

PubMed Abstract | CrossRef Full Text | Google Scholar

NIH (2002). RESEARCH ON MICROBIAL BIOFILMS. Office of Extramural Research (OER), National Institutes of Health (NIH), Department of Health and Human Services (HHS Available at: http://grants.nih.gov/grants/guide/pa-files/PA-03-047.html.

Google Scholar

Orazi G., O’Toole G. A. (2019). “It Takes a Village”: Mechanisms Underlying Antimicrobial Recalcitrance of Polymicrobial Biofilms. J. Bacteriol. 202(1):10.1128/JB.00530-19.

Google Scholar

Parlak O., Richter-Dahlfors A. (2020). Bacterial Sensing and Biofilm Monitoring for Infection Diagnostics. Macromol. Biosci. 20, e2000129. doi: 10.1002/mabi.202000129

PubMed Abstract | CrossRef Full Text | Google Scholar

Peters B. M., Jabra-Rizk M. A., O’may G. A., Costerton J. W., Shirtliff M. E. (2012). Polymicrobial interactions: impact on pathogenesis and human disease. Clin. Microbiol. Rev. 25, 193–213. doi: 10.1128/CMR.00013-11

PubMed Abstract | CrossRef Full Text | Google Scholar

Redman W. K., Welch G. S., Rumbaugh K. P. (2020). Differential Efficacy of Glycoside Hydrolases to Disperse Biofilms. Front. Cell Infect. Microbiol. 10, 379. doi: 10.3389/fcimb.2020.00379

PubMed Abstract | CrossRef Full Text | Google Scholar

Rice S. A., Mcdougald D., Kumar N., Kjelleberg S. (2005). The use of quorum-sensing blockers as therapeutic agents for the control of biofilm-associated infections. Curr. Opin. Investig. Drugs 6, 178–184.

PubMed Abstract | Google Scholar

Rode D. K. H., Singh P. K., Drescher K. (2020). Multicellular and unicellular responses of microbial biofilms to stress. Biol. Chem. 401, 1365–1374. doi: 10.1515/hsz-2020-0213

PubMed Abstract | CrossRef Full Text | Google Scholar

Røder H. L., Olsen N. M. C., Whiteley M., Burmølle M. (2020). Unravelling interspecies interactions across heterogeneities in complex biofilm communities. Environ. Microbiol. 22, 5–16. doi: 10.1111/1462-2920.14834

PubMed Abstract | CrossRef Full Text | Google Scholar

Schiessl K. T., Hu F., Jo J., Nazia S. Z., Wang B., Price-Whelan A., et al. (2019). Phenazine production promotes antibiotic tolerance and metabolic heterogeneity in Pseudomonas aeruginosa biofilms. Nat. Commun. 10, 762. doi: 10.1038/s41467-019-08733-w

PubMed Abstract | CrossRef Full Text | Google Scholar

Seneviratne C. J., Suriyanarayanan T., Widyarman A. S., Lee L. S., Lau M., Ching J., et al. (2020). Multi-omics tools for studying microbial biofilms: current perspectives and future directions. Crit. Rev. Microbiol. 46, 759–778. doi: 10.1080/1040841X.2020.1828817

PubMed Abstract | CrossRef Full Text | Google Scholar

Stewart P. S., Bjarnsholt T. (2020). Risk factors for chronic biofilm-related infection associated with implanted medical devices. Clin. Microbiol. Infect. 26, 1034–1038. doi: 10.1016/j.cmi.2020.02.027

PubMed Abstract | CrossRef Full Text | Google Scholar

Tay W. H., Chong K. K., Kline K. A. (2016). Polymicrobial-Host Interactions during Infection. J. Mol. Biol. 428, 3355–3371. doi: 10.1016/j.jmb.2016.05.006

PubMed Abstract | CrossRef Full Text | Google Scholar

Thanabalasuriar A., Scott B. N. V., Peiseler M., Willson M. E., Zeng Z., Warrener P., et al. (2019). Neutrophil Extracellular Traps Confine Pseudomonas aeruginosa Ocular Biofilms and Restrict Brain Invasion. Cell Host Microbe 25, 526–536.e524. doi: 10.1016/j.chom.2019.02.007

PubMed Abstract | CrossRef Full Text | Google Scholar

Tian S., Su L., Liu Y., Cao J., Yang G., Ren Y., et al. (2020). Self-targeting, zwitterionic micellar dispersants enhance antibiotic killing of infectious biofilms-An intravital imaging study in mice. Sci. Adv. 6, eabb1112. doi: 10.1126/sciadv.abb1112

PubMed Abstract | CrossRef Full Text | Google Scholar

Torres J. P., Senejani A. G., Gaur G., Oldakowski M., Murali K., Sapi E. (2020). Ex Vivo Murine Skin Model for B. burgdorferi Biofilm. Antibiot. (Basel) 9(9):528. doi: 10.3390/antibiotics9090528

CrossRef Full Text | Google Scholar

Valm A. M., Mark Welch J. L., Borisy G. G. (2012). CLASI-FISH: principles of combinatorial labeling and spectral imaging. Syst. Appl. Microbiol. 35, 496–502. doi: 10.1016/j.syapm.2012.03.004

PubMed Abstract | CrossRef Full Text | Google Scholar

Valm A. M. (2019). The Structure of Dental Plaque Microbial Communities in the Transition from Health to Dental Caries and Periodontal Disease. J. Mol. Biol. 431, 2957–2969. doi: 10.1016/j.jmb.2019.05.016

PubMed Abstract | CrossRef Full Text | Google Scholar

Van Dyck K., Van Dijck P., Vande Velde G. (2020). Bioluminescence Imaging to Study Mature Biofilm Formation by Candida spp. and Antifungal Activity In Vitro and In Vivo. Methods Mol. Biol. 2081, 127–143. doi: 10.1007/978-1-4939-9940-8_9

PubMed Abstract | CrossRef Full Text | Google Scholar

Whitchurch C. B., Tolker-Nielsen T., Ragas P. C., Mattick J. S. (2002). Extracellular DNA required for bacterial biofilm formation. Science 295, 1487. doi: 10.1126/science.295.5559.1487

PubMed Abstract | CrossRef Full Text | Google Scholar

Wolcott R., Costerton J. W., Raoult D., Cutler S. J. (2013). The polymicrobial nature of biofilm infection. Clin. Microbiol. Infect. 19, 107–112. doi: 10.1111/j.1469-0691.2012.04001.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Yang C. Y., Bialecka-Fornal M., Weatherwax C., Larkin J. W., Prindle A., Liu J., et al. (2020). Encoding Membrane-Potential-Based Memory within a Microbial Community. Cell Syst. 10, 417–423.e413. doi: 10.1016/j.cels.2020.04.002

PubMed Abstract | CrossRef Full Text | Google Scholar

Yuan L., De Haan P., Peterson B. W., De Jong E. D., Verpoorte E., Van Der Mei H. C., et al. (2020). Visualization of Bacterial Colonization and Cellular Layers in a Gut-on-a-Chip System Using Optical Coherence Tomography. Microsc. Microanal., 1–9. doi: 10.1017/S143192762002454X

PubMed Abstract | CrossRef Full Text | Google Scholar

Zobell C. E., Allen E. C. (1935). The Significance of Marine Bacteria in the Fouling of Submerged Surfaces. J. Bacteriol. 29, 239–251. doi: 10.1128/JB.29.3.239-251.1935

PubMed Abstract | CrossRef Full Text | Google Scholar

Keywords: biofilm infections, multi-species biofilm, infection models, biofilm detection methods, biofilm heterogeneity

Citation: Beloin C and McDougald D (2021) Speciality Grand Challenge for “Biofilms”. Front. Cell. Infect. Microbiol. 11:632429. doi: 10.3389/fcimb.2021.632429

Received: 23 November 2020; Accepted: 28 January 2021;
Published: 22 February 2021.

Edited and Reviewed by: Yousef Abu Kwaik, University of Louisville, United States

Copyright © 2021 Beloin and McDougald. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner(s) are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Christophe Beloin, cbeloin@pasteur.fr; Diane McDougald, Diane.McDougald@uts.edu.au

Download