Skip to main content
Journal of Bacteriology logoLink to Journal of Bacteriology
. 2017 Jul 11;199(15):e00012-17. doi: 10.1128/JB.00012-17

Spheres of Hope, Packets of Doom: the Good and Bad of Outer Membrane Vesicles in Interspecies and Ecological Dynamics

Jonathan B Lynch a,, Rosanna A Alegado b,
Editor: George O'Toolec
PMCID: PMC5512217  PMID: 28416709

ABSTRACT

Outer membrane vesicles (OMVs) are proteoliposome nanoparticles ubiquitously produced by Gram-negative bacteria. Typically bearing a composition similar to those of the outer membrane and periplasm of the cells from which they are derived, OMVs package an array of proteins, lipids, and nucleic acids. Once considered inconsequential by-products of bacterial growth, OMVs have since been demonstrated to mediate cellular stress relief, promote horizontal gene transfer and antimicrobial activity, and elicit metazoan inflammation. Recently, OMVs have gained appreciation as critical moderators of interorganismal dynamics. In this review, we focus on recent progress toward understanding the functions of OMVs with regard to symbiosis and ecological contexts, and we propose potential avenues for future OMV studies.

KEYWORDS: host-microbe interactions, outer membrane vesicles

INTRODUCTION

We live in a world in which chemical cues produced by bacteria influence the cellular behavior of neighboring organisms (1). Identifying the modes for material exchange between organisms will inform our understanding of how long-term relationships are established and maintained. Gram-negative bacteria use diverse mechanisms to secrete compounds, ranging from simple systems (e.g., type I and V secretion) to multiprotein complexes (type III, IV, and VI secretion) (2). However, much less attention has been devoted to the role of outer membrane vesicles (OMVs) in interspecies interactions.

OMVs are nanoparticles composed of lipid bilayers originating from the outer membrane of Gram-negative bacteria (3), ranging in diameter from 20 to 200 nm (4). These spherical bodies are produced by all Gram-negative bacteria across a broad spectrum of conditions and environments (59). OMV biogenesis is constitutive, but the production rate and composition of OMVs are sensitive to stress and environmental fluctuations (1014). First described half a century ago as membrane sacs that enable the excretion of cell wall material (15), OMVs are a generalized secretion system for complex combinations of compounds. In addition to material from the outer membrane and periplasm, OMVs frequently contain other materials in either the vesicle lumen or incorporated into the membrane bilayer, including cell wall components, nucleic acids, toxins, and labile carbon (1620). Protected within a lipid bilayer from degradation, OMV contents may persist longer in extracellular environments than exposed macromolecules released through other means.

OMV content is diverse, enabling these abiotic spheroids to have a broad range of activities. They serve as effectors, agents of genetic exchange, or shared resources and can be deployed as a defense against competing organisms or as building blocks for biofilms (7, 2124). The contributions of OMVs to host colonization and disease have been characterized (18, 2527). In contrast, the roles of OMVs in beneficial partnerships have been far less studied.

OMVs distribute materials between organisms, circumventing the need for direct cell-cell contact. In aqueous environments, these nanoparticles may migrate from the OMV producer to surprisingly distant sites within a host or ecosystem (28). For example, orally administered OMVs from Bacteroides thetaiotaomicron were capable of passing though the gut mucosal barrier in order to reach underlying epithelia and macrophages (29). Additionally, Salmonella enterica residing within the Salmonella-containing vacuole (SCV) of epithelial cells produces OMVs that escape the SCV, as well as the infected host cell, and enter neighboring cells (30). Delivery to other Gram-negative bacteria seems fairly permissive; OMVs can directly attach and fuse to the outer membrane of recipient cells (3134). After attachment, the luminal contents of the OMVs are delivered into the periplasmic space of the recipient. OMVs enter eukaryotic cells through a variety of mechanisms: direct membrane fusion (28, 34, 35), lipid rafts (36), receptor-mediated endocytosis (19, 37), and caveolin-mediated endocytosis (3842). These modes of entry potentially constrain the functions of OMVs as public goods in interdomain contexts.

Here, we evaluate the role of OMVs in mediating interspecies communication, cooperation, and conflict. We examine the benefits of OMV production for both the bacterial producer (symbiont in host contexts) and neighboring cells. Microbial symbionts compete with other symbionts and experience antagonistic host responses, and OMVs may help to reduce these burdens in mutualistic relationships, enabling microbes to attain a net benefit from their association with the host (43). We review recent findings regarding the physiological and ecological benefits of OMVs as molecular couriers in cooperative interactions and environmental settings, and we draw attention to potential lines of research on the beneficial role of OMVs.

BEING A GOOD NEIGHBOR: BENEFITS PROVIDED BY OMVs

As OMVs can be utilized by neighboring cells, they might be considered public goods (44). For instance, OMVs may facilitate cross-feeding, enabling the maintenance of complex communities in nutritionally dilute environments despite patchy resource distribution. OMVs from the cyanobacterium Prochlorococcus contain labile carbon that can be used as the sole carbon source for the heterotrophs Alteromonas and Halomonas (7). Similarly, gut-residing members of the Bacteroides genus distribute hydrolases and polysaccharide lyases via OMVs (45, 46); by making these enzymes publicly available, the microbiome can communally break down complex polysaccharides that are immutable to individual species. The exact benefit to the producer of providing these services in each of these circumstances is unclear, but it is assumed to provide a net benefit, such as a higher survival rate (44). To test this, the fitness of the OMV-producing Prochlorococcus sp. should be compared when grown alone (monoculture) or in coculture with heterotrophs from pelagic ecosystems. Specifically, varying the ratios of Prochlorococcus to consumer may more clearly illuminate the sociality of these interactions by displaying the relative effects of direct and indirect relationships between organisms.

In contrast, scavenging scarce resources has clear direct benefits for the OMV producer, as well as other “cheating” organisms. Metal ions are one such resource that OMVs may concentrate for microbial communities. Proteomic studies of OMVs from diverse bacterial species have revealed a number of metal-ion-binding proteins (47). Recent work by Lin et al. (48) elucidated a mechanism that enables Pseudomonas aeruginosa to access metal-rich OMVs via a type VI secretion system (T6SS). TseF is secreted by the T6SS and associates with the Pseudomonas quinolone signal (PQS) as well as P. aeruginosa receptors OprF and FptA. PQS has high affinity for iron, contributing to the sequestration of this ion in OMVs, which can then be utilized by P. aeruginosa under iron-limiting conditions. Thus, these proteins provide a means of scavenging and concentrating freely diffusing ions, offering a rich nutrient source to organisms able to process them. In this manner, OMVs can modulate ecosystem production and diversity within these niches.

OMVs influence interactions between members of surface-associated polymicrobial communities, as well as regulate the ability to colonize different habitats, provide structure for growth, and promote dispersal. OMVs stimulate bacterial aggregation on substrates by trafficking quorum-sensing molecules (24, 33, 49, 50). For example, Porphyromonas gingivalis OMVs promote the aggregation of other species frequently found in the oral cavity, including Staphylococcus aureus, Streptococcus, Actinomyces spp., and even the opportunistic eukaryotic pathogen Candida albicans (51). In addition to providing stimulatory cues, bacteria may secrete OMVs to enhance and reinforce the physical structure of polymicrobial biofilms (21, 22, 50). Conversely, OMVs may enable bacterial cells to transition from sessile to planktonic lifestyles, as in the case of Xylella fastidiosa OMVs that inhibit attachment to plant cells (23). Ionescu et al. hypothesize that this confers a potential dispersal benefit to bacteria. Dispersal has been proposed be a social trait, potentially by reducing competition with nondispersing relatives (52). By mediating intraspecies and interspecies social behaviors on surfaces, OMVs are key modulators within complex communities.

Finally, OMVs are likely to play a role in life history transitions in eukaryotes. As the closest living relatives of animals, a group of free-living bacterivorous microbial eukaryotes called choanoflagellates have served as a key point of comparison for understanding animal origins (5355) and the influence of bacteria on eukaryotic biology. The choanoflagellate Salpingoeca rosetta switches from a unicellular swimmer to a multicelled rosette upon exposure to lipid-based compounds produced by the marine Bacteroidetes organism Algoriphagus machipongonensis (56, 58). The morphogenic cues are highly hydrophobic sulfonolipids and are nearly insoluble in seawater, yet A. machipongonensis OMVs contain these bioactive compounds and directly fuse with unicellular S. rosetta (R. Alegado, A. Woznica, S. Cao, C. Beemelmanns, H. Turano, J. Clardy, and N. King, unpublished data). It remains unclear what benefit, if any, A. machipongonensis receives from S. rosetta as a result of this interaction; however, these data suggest that the bioactive compounds triggering multicellular development in S. rosetta may be perceived through OMVs. Thus, OMVs may be a general system for the trafficking of hydrophobic bioactive molecules, such as lipids. Indeed, OMVs are one of the only means by which bioactive lipids can be trafficked. Larval settlement and metamorphosis in several basal metazoans have been shown to require uncharacterized factors released into the water column and produced by benthic biofilms (5963), and OMVs may be the vehicle by which signals are carried. We anticipate that additional cases of OMVs exerting influence on animal biology will be discovered.

ERECTING GOOD FENCES: OMVs AS SELF-PROTECTION

OMV biogenesis is constitutive in growing cells, yet the assumed metabolic cost of producing diverse OMVs as public goods would diminish the benefits for producers (64). OMV-producing cells could limit cheaters within their ecological niche by specifically targeting competitors with destructive OMVs while directing beneficial goods to mutualistic partners, but the most likely scenario is that these goods are concurrently delivered to both friends and foes.

Notably, bactericidal OMVs can act against both Gram-positive and Gram-negative cells (31). Determining whether bacteria produce OMVs that are more effective against closely related competitors would implicate a role in niche competition (65). The composition of stress-induced OMVs indicates that they are specialized to compete for key resources. For example, Myxococcus xanthus OMVs have significantly higher concentrations of alkaline phosphatase and other lytic enzymes than the M. xanthus cell envelope, suggesting that these particles lyse cellular prey and scavenge their phosphate (66, 67).

Conversely, OMVs may also serve as a defense mechanism for producers. These abiotic particles act as decoys for phage by increasing the available membrane surface area for attachment, reducing their potential infectivity (7, 67). Infection by T4 phages was shown to increase OMV vesiculation in Escherichia coli (68), and phage PHM-2 has been shown to bind purified Prochlorococcus OMVs (7). Li et al. recently proposed that OMV-mediated phage defense may be a resilience mechanism by maintaining functional diversity of the host microbiota (24). However, the extent to which OMV decoys benefit species unrelated to the OMV-producing cell is an open question, as phage attachment is highly species specific. If it were the case that OMV phage decoys conferred a survival advantage to related bacterial species, these abiotic particles would serve as a mechanism of kin discrimination. Together, these findings suggest that bacteria produce OMVs with either positive or negative effects on closely related cells, emphasizing the complex role that OMVs play in interorganismal dynamics.

In the context of interbacterial competition and host monitoring, OMVs play key protective roles for resident microbes (6971). Supplementing E. coli or Pseudomonas syringae cultures with OMVs or a hypervesiculating mutant protected cells from a number of antimicrobial peptides that interact with negatively charged phospholipids, including polymyxin B, colistin, and melittin (67, 72). There is evidence that OMVs sequester these compounds, decreasing the effective concentration to which recipient bacterial cells are exposed. Likewise, OMVs may be capable of specifically sequestering other antagonistic molecules, such as antibodies, thereby modulating host immunity against resident microbes. OMV secretion may enhance virulence in vivo (73). Some pathogens preferentially package virulence factors in their OMVs (7375). Other pathogenic bacteria secrete more OMVs than their nonpathogenic congeners (3, 76), evidence that increased production increases individual competitive advantage.

REACHING A DÉTENTE: ROLE OF OMVs IN STABILIZING COOPERATIVE RELATIONSHIPS

The ubiquitous and versatile nature of OMVs suggests that OMVs could act at the interface of host and microbe. These particles mediate a variety of interorganismal interactions (Table 1), likely dependent on recipient cell type and OMV composition (37, 77). Furthermore, OMVs have been shown to influence a number of homeostatic processes in animals, including cell proliferation (78), apoptosis (79), and autophagy (80). Despite the physiological and evolutionary significance of mutualisms (1), there is a dearth of research on the functions of OMVs in mutualistic symbioses involving multicellular organisms.

TABLE 1.

Examples of OMV cargo with roles in intercellular interactions

Class Compound(s) (bacterial producer[s])a Recipient(s) Reference(s)
Proteinaceous Cytolethal distending toxin (Campylobacter jejuni) Human small intestine epithelial cell lines 18
Bacteroides fragilis toxin 2 (B. fragilis) Human colon epithelial cell lines 104
CFTR inhibitory factor/Cif (Pseudomonas aeruginosa) Human airway epithelial cell lines 28
OmpU (Vibrio fischeri) Euprymna scolopes phagocytic immune cells? 11
Glycoside hydrolases and polysaccharide lyases that degrade dietary polysaccharides (Bacteroides thetaiotaomicron, Bacteroides ovatus) Bacteroidales common in human intestinal microbiota (e.g., Bacteroides vulgatus) 45, 46
Ef-Tu (Xanthomonas campestris) Arabidopsis thaliana leaves 103
Carbohydrate Polysaccharide A (B. fragilis) Murine dendritic cells 29
Peptidoglycan (Helicobacter pylori) Human gastric epithelial cells 105
Nucleic acid eDNA (Pseudomonas aeruginosa) P. aeruginosa biofilms 106
Virulence-conferring DNA (Escherichia coli O157:H7) E. coli JM109, Salmonella enterica serovar Enteritidis 107
Antibiotic resistance genes (Neisseria gonorrhoeae) Penicillin-sensitive N. gonorrhoeae 108
Small RNAs (Vibrio cholerae) Unknown 109
Small RNAs (P. aeruginosa) Human bronchial epithelial cells 20
Small molecule C16-homoserine lactone (Paracoccus denitrificans) Self 33
Pseudomonas quorum signal (P. aeruginosa) Self 49
Labile carbon (Prochlorococcus marinus) Halomonas, Alteromonas 7
Lipid Rosette inducing factor 1 (Algoriphagus machipongonensis) Salpingoeca rosetta Alegado et al., unpublished
a

CFTR, cystic fibrosis transmembrane conductance regulator; eDNA, extracellular DNA.

Bacteroides fragilis is a prominent member of the human intestinal microbiota and has various capsular polysaccharides that it differentially displays. Some of these polysaccharides, such as polysaccharide A (PSA), are zwitterionic molecules that can be released into the intestinal lumen. PSA is taken up by host cells and shifts the host's immune system to a less inflammatory and regulatory T-cell-mediated state, which may ameliorate inflammatory disorders of the host (8183). Immunoeffective PSA was detected in B. fragilis OMVs, implicating them as an in vivo delivery mechanism (29). This finding highlights the potential of OMVs to positively influence host health and the need for further exploration of OMVs isolated from microbial communities from different host sites. To that end, OMVs from probiotics, such as the E. coli Nissle 1917 strain, have recently been surveyed in an attempt to scour their contents for beneficial molecules (84), and they have been found to contain factors, such as the kinase-altering protein TcpC, that can positively affect intestinal health (85, 86).

The symbiosis between the marine bacterium Vibrio fischeri and the Hawaiian bobtail squid Euprymna scolopes is mediated by compounds produced by both partners. In the juvenile squid light organ, planktonic bacteria first attach and aggregate before proceeding through a series of chemical challenges produced by the host, ultimately colonizing the lumen of the deeper crypts of the light organ for the remainder of the squid's life (87). The development of this symbiosis drives physiological responses from the squid, such as migration of blood cells and reshaping of light organ tissue to more closely resemble that of an adult (87). Recently, nonproteinaceous contents from V. fischeri OMVs were found sufficient to stimulate a subset of these developmental responses, despite a lack of compounds known to stimulate light organ development (88). Examining hyper- and hypovesiculating mutants, as well as nonvesiculators, is complicated by pleiotropic consequences on other pathways, but these mutants may still provide insight into how OMVs shape symbioses.

IMPLICATIONS OF OMVs AS MATERIAL GOODS AND MISSILES IN MICROBIAL INTERACTIONS

Experimental approaches for investigating functions of OMVs.

OMVs are constitutively released, yet they vary with environment and can promote survival for the bacterial producer (13, 67). Since OMVs benefit the producer, yet can exert a variety of effects on recipient organisms across spatial and temporal scales, they may be critical shapers of the economies between organisms and may act as currency or poison. Development of multispecies models as well as specialized host models (e.g., ex vivo studies and immunodeficient hosts) where bacteria reside in a niche with fewer host interactions will be valuable for untangling these complications. Although such models will lack a wild-type microbiota, these approaches may provide microbe-centric data that inform more natural situations.

Mechanisms underlying OMV-driven processes are difficult to unravel, due in part to the potential for OMVs to simultaneously and distinctly impact the host and other members of the microbiota. Proteomics has been frequently used to assess the functional capacity of OMVs, but additional omics approaches, particularly lipidomics, may reveal novel information about the OMV-cell interface. Recent work has demonstrated that lipid remodeling may occur prior to OMV formation (12, 13). This information can be incorporated into interaction networks that may reveal whether specific lipids are sorted into OMVs with characteristic functions as the producing cell responds to changing conditions. Synthetic liposomes that retain the chemical properties and topology of native membranes can then be employed to test biochemical necessity and sufficiency of compounds in OMV-cell interactions. OMVs from a variety of bacteria have broad effects on animal immune responses (reviewed in reference 89). Native and modified OMVs have been implemented as vaccines against pathogens, such as Salmonella enterica serovar Typhimurium and Vibrio cholerae, in animal models (26, 90, 91) and have reached clinical usage against Neisseria meningitidis (9294). Due to the size of OMVs, visualizing them in situ has generally required electron microscopy of fixed material, but improvements in superresolution techniques hold promise for tracking OMVs, their contents, and their interactions in living environments.

Accounting for OMVs in ecological community dynamics.

Gram-negative bacteria have the capacity to release prodigious amounts of OMVs, but a precise measure of how much material is being released into the extracellular milieu has not been undertaken. Recent studies demonstrating high levels of OMVs in marine environments have illuminated the global nature of these abiotic environmental factors (7). Biller et al. raised the question of how these findings may affect our calculation of nutrient budgets across the biosphere (7). With recently improved estimates of microbial abundance, we can use V. fischeri as a model to estimate the mass of certain elements locked into OMVs to approximate the global elemental impact of OMVs. A culture of V. fischeri isolate ES114 grown in rich lysogenic broth with salt (LBS) saturates around 109 CFU · ml−1 and produces approximately 10 pg of OMV protein · ml−1 of culture in 24 h, a rate of 10−17 g of protein · cell−1 · 24 h−1. Although the growth rate of V. fischeri is likely higher when grown in vitro than for bacteria in the marine environment, it enables the following back of the envelope calculation. With the estimated 9.2 × 1029 prokaryotic cells (combining Bacteria and Archaea) in Earth's oceans (95) and applying a conservative estimate of 10% Gram-negative bacteria, we estimate that roughly ∼1,000,000 metric tons of protein will be released in the ocean via OMVs every day. Assuming these proteins contain equal ratios of each of the 20 major amino acids, every 24 h, Gram-negative bacteria in the ocean release approximately 3.92 × 1011 g of carbon, 1.24 × 1011 g of nitrogen, and 1.96 × 1010 g of sulfur in OMV proteins alone. While this is a very rough estimate, it is interesting to note that it loosely matches an estimate of Prochlorococcus vesicle production (7). Although this is a small fraction compared to the oceanic reservoir of 4 × 1019 g of carbon (96) and is an overestimation of total mass being packaged into OMVs globally, even a fraction of this amount of organic material could alter marine carbon budgets in oligotrophic environments. Furthermore, the inclusion of nucleic acids, small molecules, and lipids packaged into OMVs would increase these values and present a more a sizable contribution to global elemental stores. While these estimates are based on simplistic assumptions, they illustrate the sheer potential of OMVs as a nutritional sink, or, if these materials are presented in bioaccessible forms, a nutritional bounty for organisms that can metabolize them, especially in highly competitive environments with varied access to nutrients.

Considering that isolation methods for viruses also result in enrichment of OMVs, estimates of viral abundance may be conflated with OMV abundance (97). This is a legitimate concern, as vesicles contain various amounts of DNA derived from bacterial, eukaryotic, and viral sources (98), which adds complexity to calculations describing the global impact of either group. Thus, to understand the ecological impacts of OMVs, we need to elucidate the mechanisms by which these particles are consumed, recycled, or otherwise turned over to release their molecular contents to the environment.

CONCLUDING REMARKS

OMVs have the capacity to facilitate beneficial outcomes across symbioses. In particular, OMVs shape metazoan microbiotas and promote interactions between partnered organisms by mediating physical proximity, temporal stability, and access to privileged niches within the host. On evolutionary time scales, public goods provided by OMVs may select for cooperative behavior, driving the establishment of mutualisms. Finally, we note that the preponderance of research has examined OMVs in the context of animal symbioses; to date, only a few studies have focused on OMV-plant interactions (23, 99103). Extending the role of OMVs in the biology of other multicellular lineages is a budding area for future investigation.

ACKNOWLEDGMENTS

This work was inspired by the 6th ASM Conference on Beneficial Microbes. We thank Tera Levin and our anonymous reviewers for helpful comments on the manuscript. We also thank Edward Ruby for feedback and training support to J.B.L.

This work has been supported by a National Institutes of Health National Research Service Award to J.B.L. (grant F32GM119238) and a National Science Foundation grant, IOS-1558169, to R.A.A.

We declare no conflicts of interest regarding this work.

REFERENCES

  • 1.McFall-Ngai M, Hadfield MG, Bosch TCG, Carey HV, Domazet-Loso T, Douglas AE, Dubilier N, Eberl G, Fukami T, Gilbert SF, Hentschel U, King N, Kjelleberg S, Knoll AH, Kremer N, Mazmanian SK, Metcalf JL, Nealson K, Pierce NE, Rawls JF, Reid A, Ruby EG, Rumpho M, Sanders JG, Tautz D, Wernegreen JJ. 2013. Animals in a bacterial world, a new imperative for the life sciences. Proc Natl Acad Sci U S A 110:3229–3236. doi: 10.1073/pnas.1218525110. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2.Costa TRD, Felisberto-Rodrigues C, Meir A, Prevost MS, Redzej A, Trokter M, Waksman G. 2015. Secretion systems in Gram-negative bacteria: structural and mechanistic insights. Nat Rev Microbiol 13:343–359. doi: 10.1038/nrmicro3456. [DOI] [PubMed] [Google Scholar]
  • 3.Amano A, Takeuchi H, Furuta N. 2010. Outer membrane vesicles function as offensive weapons in host-parasite interactions. Microbes Infect 12:791–798. doi: 10.1016/j.micinf.2010.05.008. [DOI] [PubMed] [Google Scholar]
  • 4.Bonnington KE, Kuehn MJ. 2014. Protein selection and export via outer membrane vesicles. Biochim Biophys Acta 1843:1612–1619. doi: 10.1016/j.bbamcr.2013.12.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Beveridge TJ. 1999. Structures of Gram-negative cell walls and their derived membrane vesicles. J Bacteriol 181:4725–4733. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Beveridge TJ, Makin SA, Kadurugamuwa JL, Li Z. 1997. Interactions between biofilms and the environment. FEMS Microbiol Rev 20:291–303. doi: 10.1111/j.1574-6976.1997.tb00315.x. [DOI] [PubMed] [Google Scholar]
  • 7.Biller SJ, Schubotz F, Roggensack SE, Thompson AW, Summons RE, Chisholm SW. 2014. Bacterial vesicles in marine ecosystems. Science 343:183–186. doi: 10.1126/science.1243457. [DOI] [PubMed] [Google Scholar]
  • 8.Hickey CA, Kuhn KA, Donermeyer DL, Porter NT, Jin C, Cameron EA, Jung H, Kaiko GE, Wegorzewska M, Malvin NP, Glowacki RWP, Hansson GC, Allen PM, Martens EC, Stappenbeck TS. 2015. Colitogenic Bacteroides thetaiotaomicron antigens access host immune cells in a sulfatase-dependent manner via outer membrane vesicles. Cell Host Microbe 17:672–680. doi: 10.1016/j.chom.2015.04.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Hellman J, Loiselle PM, Zanzot EM, Allaire JE, Tehan MM, Boyle LA, Kurnick JT, Warren HS. 2000. Release of Gram-negative outer-membrane proteins into human serum and septic rat blood and their interactions with immunoglobulin in antiserum to Escherichia coli J5. J Infect Dis 181:1034–1043. doi: 10.1086/315302. [DOI] [PubMed] [Google Scholar]
  • 10.Orench-Rivera N, Kuehn MJ. 2016. Environmentally controlled bacterial vesicle-mediated export. Cell Microbiol 18:1525–1536. doi: 10.1111/cmi.12676. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Aschtgen M-S, Lynch JB, Koch E, Schwartzman J, McFall-Ngai M, Ruby E. 2016. Rotation of Vibrio fischeri flagella produces outer membrane vesicles that induce host development. J Bacteriol 198:2156–2165. doi: 10.1128/JB.00101-16. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Bonnington KE, Kuehn MJ. 2016. Outer membrane vesicle production facilitates LPS remodeling and outer membrane maintenance in Salmonella during environmental transitions. mBio 7(5):e01532-16. doi: 10.1128/mBio.01532-16. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Elhenawy W, Bording-Jorgensen M, Valguarnera E, Haurat MF, Wine E, Feldman MF. 2016. LPS remodeling triggers formation of outer membrane vesicles in Salmonella. mBio 7(4):e00940-16. doi: 10.1128/mBio.00940-16. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Roier S, Zingl FG, Cakar F, Durakovic S, Kohl P, Eichmann TO, Klug L, Gadermaier B, Weinzerl K, Prassl R, Lass A, Daum G, Reidl J, Feldman MF, Schild S. 2016. A novel mechanism for the biogenesis of outer membrane vesicles in Gram-negative bacteria. Nat Commun 7:10515. doi: 10.1038/ncomms10515. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Chatterjee SN, Das J. 1967. Electron microscopic observations on the excretion of cell-wall material by Vibrio cholerae. Microbiology 49:1–11. [DOI] [PubMed] [Google Scholar]
  • 16.Renelli M, Matias V, Lo RY, Beveridge TJ. 2004. DNA-containing membrane vesicles of Pseudomonas aeruginosa PAO1 and their genetic transformation potential. Microbiology 150:2161–2169. doi: 10.1099/mic.0.26841-0. [DOI] [PubMed] [Google Scholar]
  • 17.Kaparakis M, Turnbull L, Carneiro L, Firth S, Coleman HA, Parkington HC, Le Bourhis L, Karrar A, Viala J, Mak J, Hutton ML, Davies JK, Crack PJ, Hertzog PJ, Philpott DJ, Girardin SE, Whitchurch CB, Ferrero RL. 2010. Bacterial membrane vesicles deliver peptidoglycan to NOD1 in epithelial cells. Cell Microbiol 12:372–385. doi: 10.1111/j.1462-5822.2009.01404.x. [DOI] [PubMed] [Google Scholar]
  • 18.Lindmark B, Rompikuntal PK, Vaitkevicius K, Song T, Mizunoe Y, Uhlin BE, Guerry P, Wai SN. 2009. Outer membrane vesicle-mediated release of cytolethal distending toxin (CDT) from Campylobacter jejuni. BMC Microbiol 9:220–229. doi: 10.1186/1471-2180-9-220. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Vanaja SK, Russo AJ, Behl B, Banerjee I, Yankova M, Deshmukh SD, Rathinam VA. 2016. Bacterial outer membrane vesicles mediate cytosolic localization of LPS and caspase-11 activation. Cell 165:1106–1119. doi: 10.1016/j.cell.2016.04.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Koeppen K, Hampton TH, Jarek M, Scharfe M, Gerber SA, Mielcarz DW, Demers EG, Dolben EL, Hammond JH, Hogan DA, Stanton BA. 2016. A novel mechanism of host-pathogen interaction through sRNA in bacterial outer membrane vesicles. PLoS Pathog 12:e1005672. doi: 10.1371/journal.ppat.1005672. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Kulp A, Kuehn MJ. 2010. Biological functions and biogenesis of secreted bacterial outer membrane vesicles. Annu Rev Microbiol 64:163–184. doi: 10.1146/annurev.micro.091208.073413. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Berleman J, Auer M. 2013. The role of bacterial outer membrane vesicles for intra- and interspecies delivery. Environ Microbiol 15:347–354. doi: 10.1111/1462-2920.12048. [DOI] [PubMed] [Google Scholar]
  • 23.Ionescu M, Zaini PA, Baccari C, Tran S, da Silva AM, Lindow SE. 2014. Xylella fastidiosa outer membrane vesicles modulate plant colonization by blocking attachment to surfaces. Proc Natl Acad Sci U S A 111:E3910–E3918. doi: 10.1073/pnas.1414944111. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Li J, Azam F, Zhang S. 2016. Outer membrane vesicles containing signalling molecules and active hydrolytic enzymes released by a coral pathogen Vibrio shilonii AK1. Environ Microbiol 18:3850–3866. doi: 10.1111/1462-2920.13344. [DOI] [PubMed] [Google Scholar]
  • 25.Roy K, Hamilton DJ, Munson GP, Fleckenstein JM. 2011. Outer membrane vesicles induce immune responses to virulence proteins and protect against colonization by enterotoxigenic Escherichia coli. Clin Vaccine Immunol 18:1803–1808. doi: 10.1128/CVI.05217-11. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Alaniz RC, Deatherage BL, Lara JC, Cookson BT. 2007. Membrane vesicles are immunogenic facsimiles of Salmonella Typhimurium that potently activate dendritic cells, prime B and T cell responses, and stimulate protective immunity in vivo. J Immunol 179:7692–7701. doi: 10.4049/jimmunol.179.11.7692. [DOI] [PubMed] [Google Scholar]
  • 27.Kuehn MJ, Kesty NC. 2005. Bacterial outer membrane vesicles and the host-pathogen interaction. Genes Dev 19:2645–2655. doi: 10.1101/gad.1299905. [DOI] [PubMed] [Google Scholar]
  • 28.Bomberger JM, Maceachran DP, Coutermarsh BA, Ye S, O'Toole GA, Stanton BA. 2009. Long-distance delivery of bacterial virulence factors by Pseudomonas aeruginosa outer membrane vesicles. PLoS Pathog 5:e1000382. doi: 10.1371/journal.ppat.1000382. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Shen Y, Giardino Torchia ML, Lawson GW, Karp CL, Ashwell JD, Mazmanian SK. 2012. Outer membrane vesicles of a human commensal mediate immune regulation and disease protection. Cell Host Microbe 12:509–520. doi: 10.1016/j.chom.2012.08.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Guidi R, Levi L, Rouf SF, Puiac S, Rhen M, Frisan T. 2013. Salmonella enterica delivers its genotoxin through outer membrane vesicles secreted from infected cells. Cell Microbiol 15:2034–2050. doi: 10.1111/cmi.12172. [DOI] [PubMed] [Google Scholar]
  • 31.Kadurugamuwa JL, Beveridge TJ. 1996. Bacteriolytic effect of membrane vesicles from Pseudomonas aeruginosa on other bacteria including pathogens: conceptually new antibiotics. J Bacteriol 178:2767–2774. doi: 10.1128/jb.178.10.2767-2774.1996. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Kadurugamuwa JL, Beveridge TJ. 1999. Membrane vesicles derived from Pseudomonas aeruginosa and Shigella flexneri can be integrated into the surfaces of other Gram-negative bacteria. Microbiology 145:2051–2060. doi: 10.1099/13500872-145-8-2051. [DOI] [PubMed] [Google Scholar]
  • 33.Toyofuku M, Morinaga K, Hashimoto Y, Uhl J, Shimamura H, Inaba H, Schmitt-Kopplin P, Eberl L, Nomura N. 2017. Membrane vesicle-mediated bacterial communication. ISME J 181:1203. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Jäger J, Keese S, Roessle M, Steinert M, Schromm AB. 2015. Fusion of Legionella pneumophila outer membrane vesicles with eukaryotic membrane systems is a mechanism to deliver pathogen factors to host cell membranes. Cell Microbiol 17:607–620. doi: 10.1111/cmi.12392. [DOI] [PubMed] [Google Scholar]
  • 35.Crowley JT, Toledo AM, Larocca TJ, Coleman JL, London E, Benach JL. 2013. Lipid exchange between Borrelia burgdorferi and host cells. PLoS Pathog 9:e1003109. doi: 10.1371/journal.ppat.1003109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Furuta N, Tsuda K, Omori H, Yoshimori T, Yoshimura F, Amano A. 2009. Porphyromonas gingivalis outer membrane vesicles enter human epithelial cells via an endocytic pathway and are sorted to lysosomal compartments. Infect Immun 77:4187–4196. doi: 10.1128/IAI.00009-09. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Parker H, Chitcholtan K, Hampton MB, Keenan JI. 2010. Uptake of Helicobacter pylori outer membrane vesicles by gastric epithelial cells. Infect Immun 78:5054–5061. doi: 10.1128/IAI.00299-10. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Sharpe SW, Kuehn MJ, Mason KM. 2011. Elicitation of epithelial cell-derived immune effectors by outer membrane vesicles of nontypeable Haemophilus influenzae. Infect Immun 79:4361–4369. doi: 10.1128/IAI.05332-11. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Elmi A, Watson E, Sandu P, Gundogdu O, Mills DC, Inglis NF, Manson E, Imrie L, Bajaj-Elliott M, Wren BW, Smith DGE, Dorrell N. 2012. Campylobacter jejuni outer membrane vesicles play an important role in bacterial interactions with human intestinal epithelial cells. Infect Immun 80:4089–4098. doi: 10.1128/IAI.00161-12. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Thay B, Damm A, Kufer TA, Wai SN, Oscarsson J. 2014. Aggregatibacter actinomycetemcomitans outer membrane vesicles are internalized in human host cells and trigger NOD1- and NOD2-dependent NF-κB activation. Infect Immun 82:4034–4046. doi: 10.1128/IAI.01980-14. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Mondal A, Tapader R, Chatterjee NS, Ghosh A, Sinha R, Koley H, Saha DR, Chakrabarti MK, Wai SN, Pal A. 2016. Cytotoxic and inflammatory responses induced by outer membrane vesicle-associated biologically active proteases from Vibrio cholerae. Infect Immun 84:1478–1490. doi: 10.1128/IAI.01365-15. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Kim YR, Kim BU, Kim SY, Kim CM, Na HS, Koh JT, Choy HE, Rhee JH, Lee SE. 2010. Outer membrane vesicles of Vibrio vulnificus deliver cytolysin-hemolysin VvhA into epithelial cells to induce cytotoxicity. Biochem Biophys Res Commun 399:607–612. doi: 10.1016/j.bbrc.2010.07.122. [DOI] [PubMed] [Google Scholar]
  • 43.Garcia JR, Gerardo NM. 2014. The symbiont side of symbiosis: do microbes really benefit? Front Microbiol 5:510. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.West SA, Diggle SP, Buckling A, Gardner A, Griffin AS. 2007. The social lives of microbes. Annu Rev Ecol Syst 38:53–77. doi: 10.1146/annurev.ecolsys.38.091206.095740. [DOI] [Google Scholar]
  • 45.Rakoff-Nahoum S, Coyne MJ, Comstock LE. 2014. An ecological network of polysaccharide utilization among human intestinal symbionts. Curr Biol 24:40–49. doi: 10.1016/j.cub.2013.10.077. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Elhenawy W, Debelyy MO, Feldman MF. 2014. Preferential packing of acidic glycosidases and proteases into Bacteroides outer membrane vesicles. mBio 5(2):e00909-14. doi: 10.1128/mBio.00909-14. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Lee E-Y, Bang JY, Park GW, Choi D-S, Kang JS, Kim H-J, Park K-S, Lee J-O, Kim Y-K, Kwon K-H, Kim K-P, Gho YS. 2007. Global proteomic profiling of native outer membrane vesicles derived from Escherichia coli. Proteomics 7:3143–3153. doi: 10.1002/pmic.200700196. [DOI] [PubMed] [Google Scholar]
  • 48.Lin J, Zhang W, Cheng J, Yang X, Zhu K, Wang Y, Wei G, Qian P-Y, Luo Z-Q, Shen X. 2017. A Pseudomonas T6SS effector recruits PQS-containing outer membrane vesicles for iron acquisition. Nat Commun 8:1–12. doi: 10.1038/s41467-016-0009-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Mashburn LM, Whiteley M. 2005. Membrane vesicles traffic signals and facilitate group activities in a prokaryote. Nature 437:422–425. doi: 10.1038/nature03925. [DOI] [PubMed] [Google Scholar]
  • 50.Schooling SR, Beveridge TJ. 2006. Membrane vesicles: an overlooked component of the matrices of biofilms. J Bacteriol 188:5945–5957. doi: 10.1128/JB.00257-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Kamaguchi A, Nakayama K, Ichiyama S, Nakamura R, Watanabe T, Ohta M, Baba H, Ohyama T. 2003. Effect of Porphyromonas gingivalis vesicles on coaggregation of Staphylococcus aureus to oral microorganisms. Curr Microbiol 47:485–491. doi: 10.1007/s00284-003-4069-6. [DOI] [PubMed] [Google Scholar]
  • 52.Hamilton WD, May RM. 1977. Dispersal in stable habitats. Nature 269:578–581. doi: 10.1038/269578a0. [DOI] [Google Scholar]
  • 53.King N, Hittinger CT, Carroll SB. 2003. Evolution of key cell signaling and adhesion protein families predates animal origins. Science 301:361–363. doi: 10.1126/science.1083853. [DOI] [PubMed] [Google Scholar]
  • 54.King N, Carroll SB. 2001. A receptor tyrosine kinase from choanoflagellates: molecular insights into early animal evolution. Proc Natl Acad Sci U S A 98:15032–15037. doi: 10.1073/pnas.261477698. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.King N, Westbrook MJ, Young SL, Kuo A, Abedin M, Chapman J, Fairclough S, Hellsten U, Isogai Y, Letunic I, Marr M, Pincus D, Putnam N, Rokas A, Wright KJ, Zuzow R, Dirks W, Good M, Goodstein D, Lemons D, Li W, Lyons JB, Morris A, Nichols SA, Richter DJ, Salamov A, Sequencing JGI, Bork P, Lim WA, Manning G, Miller WT, Mcginnis W, Shapiro H, Tjian R, Grigoriev IV, Rokhsar D. 2008. The genome of the choanoflagellate Monosiga brevicollis and the origin of metazoans. Nature 451:783–788. doi: 10.1038/nature06617. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Alegado RA, Brown LW, Cao S, Dermenjian RK, Zuzow R, Fairclough SR, Clardy J, King N. 2012. A bacterial sulfonolipid triggers multicellular development in the closest living relatives of animals. eLife 1:e00013. doi: 10.7554/eLife.00013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Reference deleted.
  • 58.Woznica A, Cantley AM, Beemelmanns C, Freinkman E, Clardy J, King N. 2016. Bacterial lipids activate, synergize, and inhibit a developmental switch in choanoflagellates. Proc Natl Acad Sci U S A 113:7894–7899. doi: 10.1073/pnas.1605015113. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Woollacott RM, Hadfield MG. 1996. Induction of metamorphosis in larvae of a sponge. Invert Biol 115:257–262. doi: 10.2307/3227015. [DOI] [Google Scholar]
  • 60.Webster NS, Smith LD, Heyward AJ, Watts JEM, Webb RI, Blackall LL, Negri AP. 2004. Metamorphosis of a scleractinian coral in response to microbial biofilms. Appl Environ Microbiol 70:1213–1221. doi: 10.1128/AEM.70.2.1213-1221.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Bertrand JF, Woollacott RM. 2003. G protein-linked receptors and induction of metamorphosis in Bugula stolonifera (Bryozoa). Invert Biol 122:380–385. doi: 10.1111/j.1744-7410.2003.tb00102.x. [DOI] [Google Scholar]
  • 62.Wieczorek SK, Todd CD. 1997. Inhibition and facilitation of bryozoan and ascidian settlement by natural multi-species biofilms: effects of film age and the roles of active and passive larval attachment. Mar Biol 128:463–473. doi: 10.1007/s002270050113. [DOI] [Google Scholar]
  • 63.Hadfield MG. 2011. Biofilms and marine invertebrate larvae: what bacteria produce that larvae use to choose settlement sites. Annu Rev Mar Sci 3:453–470. doi: 10.1146/annurev-marine-120709-142753. [DOI] [PubMed] [Google Scholar]
  • 64.Schwartz MW, Hoeksema JD. 1998. Specialization and resource trade: biological markets as a model of mutualisms. Ecology 79:1029–1038. doi: 10.1890/0012-9658(1998)079[1029:SARTBM]2.0.CO;2. [DOI] [Google Scholar]
  • 65.Li Z, Clarke AJ, Beveridge TJ. 1998. Gram-negative bacteria produce membrane vesicles which are capable of killing other bacteria. J Bacteriol 180:5478–5483. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Evans AGL, Davey HM, Cookson A, Currinn H, Cooke-Fox G, Stanczyk PJ, Whitworth DE. 2012. Predatory activity of Myxococcus xanthus outer-membrane vesicles and properties of their hydrolase cargo. Microbiology 158:2742–2752. doi: 10.1099/mic.0.060343-0. [DOI] [PubMed] [Google Scholar]
  • 67.Manning AJ, Kuehn MJ. 2011. Contribution of bacterial outer membrane vesicles to innate bacterial defense. BMC Microbiol 11:258. doi: 10.1186/1471-2180-11-258. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Loeb MR, Kilner J. 1978. Release of a special fraction of the outer membrane from both growing and phage T4-infected Escherichia coli B. Biochim Biophys Acta 514:117–127. doi: 10.1016/0005-2736(78)90081-0. [DOI] [PubMed] [Google Scholar]
  • 69.Ellis TN, Leiman SA, Kuehn MJ. 2010. Naturally produced outer membrane vesicles from Pseudomonas aeruginosa elicit a potent innate immune response via combined sensing of both lipopolysaccharide and protein components. Infect Immun 78:3822–3831. doi: 10.1128/IAI.00433-10. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Perez Vidakovics MLA, Jendholm J, Mörgelin M, Månsson A, Larsson C, Cardell L-O, Riesbeck K. 2010. B cell activation by outer membrane vesicles—a novel virulence mechanism. PLoS Pathog 6:e1000724. doi: 10.1371/journal.ppat.1000724. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Chattopadhyay MK, Jaganandham MV. 2015. Vesicles-mediated resistance to antibiotics in bacteria. Front Microbiol 6:758. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Kulkarni HM, Swamy CV, Jagannadham MV. 2014. Molecular characterization and functional analysis of outer membrane vesicles from the antarctic bacterium Pseudomonas syringae suggest a possible response to environmental conditions. J Proteome Res 13:1345–1358. [DOI] [PubMed] [Google Scholar]
  • 73.Horstman AL, Kuehn MJ. 2000. Enterotoxigenic Escherichia coli secretes active heat-labile enterotoxin via outer membrane vesicles. J Biol Chem 275:12489–12496. doi: 10.1074/jbc.275.17.12489. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74.Wai SN, Lindmark B, Söderblom T, Takade A, Westermark M, Oscarsson J, Jass J, Richter-Dahlfors A, Mizunoe Y, Uhlin BE. 2003. Vesicle-mediated export and assembly of pore-forming oligomers of the enterobacterial ClyA cytotoxin. Cell 115:25–35. doi: 10.1016/S0092-8674(03)00754-2. [DOI] [PubMed] [Google Scholar]
  • 75.Ricci V, Chiozzi V, Necchi V, Oldani A, Romano M, Solcia E, Ventura U. 2005. Free-soluble and outer membrane vesicle-associated VacA from Helicobacter pylori: two forms of release, a different activity. Biochem Biophys Res Commun 337:173–178. doi: 10.1016/j.bbrc.2005.09.035. [DOI] [PubMed] [Google Scholar]
  • 76.Bauman SJ, Kuehn MJ. 2009. Pseudomonas aeruginosa vesicles associate with and are internalized by human lung epithelial cells. BMC Microbiol 9:26. doi: 10.1186/1471-2180-9-26. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.Pollak CN, Delpino MV, Fossati CA, Baldi PC. 2012. Outer membrane vesicles from Brucella abortus promote bacterial internalization by human monocytes and modulate their innate immune response. PLoS One 7:e50214. doi: 10.1371/journal.pone.0050214. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78.Jones C, Sadarangani M, Lewis S, Payne I, Saleem M, Derrick JP, Pollard AJ. 2016. Characterisation of the immunomodulatory effects of meningococcal Opa proteins on human peripheral blood mononuclear cells and CD4+ T cells. PLoS One 11:e0154153. doi: 10.1371/journal.pone.0154153. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 79.Kunsmann L, Rüter C, Bauwens A, Greune L, Glüder M, Kemper B, Fruth A, Wai SN, He X, Lloubes R, Schmidt MA, Dobrindt U, Mellmann A, Karch H, Bielaszewska M. 2015. Virulence from vesicles: novel mechanisms of host cell injury by Escherichia coli O104:H4 outbreak strain. Sci Rep 5:13252. doi: 10.1038/srep13252. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80.Irving AT, Mimuro H, Kufer TA, Lo C, Wheeler R, Turner LJ, Thomas BJ, Malosse C, Gantier MP, Casillas LN, Votta BJ, Bertin J, Boneca IG, Sasakawa C, Philpott DJ, Ferrero RL, Kaparakis-Liaskos M. 2014. The immune receptor NOD1 and kinase RIP2 interact with bacterial peptidoglycan on early endosomes to promote autophagy and inflammatory signaling. Cell Host Microbe 15:623–635. doi: 10.1016/j.chom.2014.04.001. [DOI] [PubMed] [Google Scholar]
  • 81.Mazmanian SK, Liu CH, Tzianabos AO, Kasper DL. 2005. An immunomodulatory molecule of symbiotic bacteria directs maturation of the host immune system. Cell 122:107–118. doi: 10.1016/j.cell.2005.05.007. [DOI] [PubMed] [Google Scholar]
  • 82.Mazmanian SK, Round JL, Kasper DL. 2008. A microbial symbiosis factor prevents intestinal inflammatory disease. Nature 453:620–625. doi: 10.1038/nature07008. [DOI] [PubMed] [Google Scholar]
  • 83.Chu H, Khosravi A, Kusumawardhani IP, Kwon AHK, Vasconcelos AC, Cunha LD, Mayer AE, Shen Y, Wu W-L, Kambal A, Targan SR, Xavier RJ, Ernst PB, Green DR, McGovern DPB, Virgin HW, Mazmanian SK. 2016. Gene-microbiota interactions contribute to the pathogenesis of inflammatory bowel disease. Science 352:1116–1120. doi: 10.1126/science.aad9948. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 84.Aguilera L, Toloza L, Giménez R, Odena A, Oliveira E, Aguilar J, Badia J, Baldomà L. 2014. Proteomic analysis of outer membrane vesicles from the probiotic strain Escherichia coli Nissle 1917. Proteomics 14:222–229. doi: 10.1002/pmic.201300328. [DOI] [PubMed] [Google Scholar]
  • 85.Tettmann B, Niewerth C, Kirschhöfer F, Neidig A, Dötsch A, Brenner-Weiss G, Fetzner S, Overhage J. 2016. Enzyme-mediated quenching of the Pseudomonas quinolone signal (PQS) promotes biofilm formation of Pseudomonas aeruginosa by increasing iron availability. Front Microbiol 7:1978. doi: 10.3389/fmicb.2016.01978. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86.Hering NA, Richter JF, Fromm A, Wieser A, Hartmann S, Günzel D, Bücker R, Fromm M, Schulzke JD, Troeger H. 2014. TcpC protein from E. coli Nissle improves epithelial barrier function involving PKCζ and ERK1/2 signaling in HT-29/B6 cells. Mucosal Immunol 7:369–378. doi: 10.1038/mi.2013.55. [DOI] [PubMed] [Google Scholar]
  • 87.McFall-Ngai MJ. 2014. The importance of microbes in animal development: lessons from the squid-Vibrio symbiosis. Annu Rev Microbiol 68:177–194. doi: 10.1146/annurev-micro-091313-103654. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 88.Aschtgen M-S, Wetzel K, Goldman W, McFall-Ngai M, Ruby E. 2016. Vibrio fischeri-derived outer membrane vesicles trigger host development. Cell Microbiol 18:488–499. doi: 10.1111/cmi.12525. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 89.Kaparakis-Liaskos M, Ferrero RL. 2015. Immune modulation by bacterial outer membrane vesicles. Nat Rev Immunol 15:375–387. doi: 10.1038/nri3837. [DOI] [PubMed] [Google Scholar]
  • 90.Roberts R, Moreno G, Bottero D, Gaillard ME, Fingermann M, Graieb A, Rumbo M, Hozbor D. 2008. Outer membrane vesicles as acellular vaccine against pertussis. Vaccine 26:4639–4646. doi: 10.1016/j.vaccine.2008.07.004. [DOI] [PubMed] [Google Scholar]
  • 91.Ormazábal M, Bartel E, Gaillard ME, Bottero D, Errea A, Zurita ME, Moreno G, Rumbo M, Castuma C, Flores D, Martín MJ, Hozbor D. 2014. Characterization of the key antigenic components of pertussis vaccine based on outer membrane vesicles. Vaccine 32:6084–6090. doi: 10.1016/j.vaccine.2014.08.084. [DOI] [PubMed] [Google Scholar]
  • 92.Mowlaboccus S, Perkins TT, Smith H, Sloots T, Tozer S, Prempeh L-J, Tay CY, Peters F, Speers D, Keil AD, Kahler CM. 2016. Temporal changes in BEXSERO antigen sequence type associated with genetic lineages of Neisseria meningitidis over a 15-year period in Western Australia. PLoS One 11:e0158315. doi: 10.1371/journal.pone.0158315. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 93.Pajon R, Lujan E, Granoff DM. 2016. A meningococcal NOMV-FHbp vaccine for Africa elicits broader serum bactericidal antibody responses against serogroup B and non-B strains than a licensed serogroup B vaccine. Vaccine 34:643–649. doi: 10.1016/j.vaccine.2015.12.034. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94.Dowling DJ, Sanders H, Cheng WK, Joshi S, Brightman S, Bergelson I, Pietrasanta C, van Haren SD, van Amsterdam S, Fernandez J, van den Dobbelsteen GPJM, Levy O. 2016. A meningococcal outer membrane vesicle vaccine incorporating genetically attenuated endotoxin dissociates inflammation from immunogenicity. Front Immunol 7:562. doi: 10.3389/fimmu.2016.00562. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 95.Kallmeyer J, Pockalny R, Adhikari RR, Smith DC, D'Hondt S. 2012. Global distribution of microbial abundance and biomass in subseafloor sediment. Proc Natl Acad Sci U S A 109:16213–16216. doi: 10.1073/pnas.1203849109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 96.Bowler C, Karl DM, Colwell RR. 2009. Microbial oceanography in a sea of opportunity. Nature 459:180–184. doi: 10.1038/nature08056. [DOI] [PubMed] [Google Scholar]
  • 97.Soler N, Krupovic M, Marguet E, Forterre P. 2015. Membrane vesicles in natural environments: a major challenge in viral ecology. ISME J 9:793–796. doi: 10.1038/ismej.2014.184. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98.Biller SJ, McDaniel LD, Breitbart M, Rogers E, Paul JH, Chisholm SW. 2017. Membrane vesicles in sea water: heterogeneous DNA content and implications for viral abundance estimates. ISME J 11:394–404. doi: 10.1038/ismej.2016.134. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99.Nascimento R, Gouran H, Chakraborty S, Gillespie HW, Almeida-Souza HO, Tu A, Rao BJ, Feldstein PA, Bruening G, Goulart LR, Dandekar AM. 2016. The type II secreted lipase/esterase LesA is a key virulence factor required for Xylella fastidiosa pathogenesis in grapevines. Sci Rep 6:18598. doi: 10.1038/srep18598. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100.Solé M, Scheibner F, Hoffmeister A-K, Hartmann N, Hause G, Rother A, Jordan M, Lautier M, Arlat M, Büttner D. 2015. Xanthomonas campestris pv. vesicatoria secretes proteases and xylanases via the Xps type II secretion system and outer membrane vesicles. J Bacteriol 197:2879–2893. doi: 10.1128/JB.00322-15. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101.Chowdhury C, Jagannadham MV. 2013. Virulence factors are released in association with outer membrane vesicles of Pseudomonas syringae pv. tomato T1 during normal growth. Biochim Biophys Acta 1834:231–239. doi: 10.1016/j.bbapap.2012.09.015. [DOI] [PubMed] [Google Scholar]
  • 102.Sidhu VK, Vorhölter F-J, Niehaus K, Watt SA. 2008. Analysis of outer membrane vesicle associated proteins isolated from the plant pathogenic bacterium Xanthomonas campestris pv. campestris. BMC Microbiol 8:87. doi: 10.1186/1471-2180-8-87. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 103.Bahar O, Mordukhovich G, Luu DD, Schwessinger B, Daudi A, Jehle AK, Felix G, Ronald PC. 2016. Bacterial outer membrane vesicles induce plant immune responses. Mol Plant Microbe Interact 29:374–384. doi: 10.1094/MPMI-12-15-0270-R. [DOI] [PubMed] [Google Scholar]
  • 104.Zakharzhevskaya NB, Tsvetkov VB, Vanyushkina AA, Varizhuk AM, Rakitina DV, Podgorsky VV, Vishnyakov IE, Kharlampieva DD, Manuvera VA, Lisitsyn FV, Gushina EA, Lazarev VN, Govorun VM. 2017. Interaction of Bacteroides fragilis toxin with outer membrane vesicles reveals new mechanism of its secretion and delivery. Front Cell Infect Microbiol 7:2. doi: 10.3389/fcimb.2017.00002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 105.Cecil JD, O'Brien-Simpson NM, Lenzo JC, Holden JA, Chen Y-Y, Singleton W, Gause KT, Yan Y, Caruso F, Reynolds EC. 2016. Differential responses of pattern recognition receptors to outer membrane vesicles of three periodontal pathogens. PLoS One 11:e0151967. doi: 10.1371/journal.pone.0151967. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 106.Turnbull L, Toyofuku M, Hynen AL, Kurosawa M, Pessi G, Petty NK, Osvath SR, Carcamo-Oyarce G, Gloag ES, Shimoni R, Omasits U, Yap X, Ito S, Cavaliere R, Monahan LG, Charles IG, Ahrens CH, Nomura N, Eberl L, Whitchurch CB. 2016. Explosive cell lysis as a mechanism for the biogenesis of bacterial membrane vesicles and biofilms. Nat Commun 7:11220. doi: 10.1038/ncomms11220. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 107.Yaron S, Kolling GL, Simon L, Matthews KR. 2000. Vesicle-mediated transfer of virulence genes from Escherichia coli O157:H7 to other enteric bacteria. Appl Environ Microbiol 66:4414–4420. doi: 10.1128/AEM.66.10.4414-4420.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 108.Dorward DW, Garon CF, Judd RC. 1989. Export and intercellular transfer of DNA via membrane blebs of Neisseria gonorrhoeae. J Bacteriol 171:2499–2505. doi: 10.1128/jb.171.5.2499-2505.1989. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 109.Sjöström AE, Sandblad L, Uhlin BE, Wai SN. 2015. Membrane vesicle-mediated release of bacterial RNA. Sci Rep 1–10. [DOI] [PMC free article] [PubMed] [Google Scholar]

Articles from Journal of Bacteriology are provided here courtesy of American Society for Microbiology (ASM)

RESOURCES