首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
Biofilms are considered to be highly resistant to antimicrobial agents. Several mechanisms have been proposed to explain this high resistance of biofilms, including restricted penetration of antimicrobial agents into biofilms, slow growth owing to nutrient limitation, expression of genes involved in the general stress response, and emergence of a biofilm-specific phenotype. However, since combinations of these factors are involved in most biofilm studies, it is still difficult to fully understand the mechanisms of biofilm resistance to antibiotics. In this study, the antibiotic susceptibility of Escherichia coli cells in biofilms was investigated with exclusion of the effects of the restricted penetration of antimicrobial agents into biofilms and the slow growth owing to nutrient limitation. Three different antibiotics, ampicillin (100 μg/ml), kanamycin (25 μg/ml), and ofloxacin (10 μg/ml), were applied directly to cells in the deeper layers of mature biofilms that developed in flow cells after removal of the surface layers of the biofilms. The results of the antibiotic treatment analyses revealed that ofloxacin and kanamycin were effective against biofilm cells, whereas ampicillin did not kill the cells, resulting in regrowth of the biofilm after the ampicillin treatment was discontinued. LIVE/DEAD staining revealed that a small fraction of resistant cells emerged in the deeper layers of the mature biofilms and that these cells were still alive even after 24 h of ampicillin treatment. Furthermore, to determine which genes in the biofilm cells are induced, allowing increased resistance to ampicillin, global gene expression was analyzed at different stages of biofilm formation, the attachment, colony formation, and maturation stages. The results showed that significant changes in gene expression occurred during biofilm formation, which were partly induced by rpoS expression. Based on the experimental data, it is likely that the observed resistance of biofilms can be attributed to formation of ampicillin-resistant subpopulations in the deeper layers of mature biofilms but not in young colony biofilms and that the production and resistance of the subpopulations were aided by biofilm-specific phenotypes, like slow growth and induction of rpoS-mediated stress responses.Reduced susceptibility of biofilm bacteria to antimicrobial agents is a crucial problem for treatment of chronic infections (11, 29, 48). It has been estimated that 65% of microbial infections are associated with biofilms (11, 29, 37), and biofilm cells are 100 to 1,000 times more resistant to antimicrobial agents than planktonic bacterial cells (11, 29, 32).The molecular nature of this apparent resistance has not been elucidated well, and a number of mechanisms have been proposed to explain the reduced susceptibility, such as restricted antibiotic penetration (47), decreased growth rates and metabolism (7, 52), quorum sensing and induction of a biofilm-specific phenotype (8, 29, 35, 39, 49), stress response activation (7, 52), and an increase in expression of efflux pumps (14). Biofilm resistance has generally been assumed to be due to the fact that the cells in the deeper layers of thick biofilms, which grow more slowly, have less access to antibiotics and nutrients. However, this is not the only reason in many cases. Familiar mechanisms of antibiotic resistance, such as modifying enzymes and target mutations, do not seem to be responsible for the biofilm resistance. Even sensitive bacteria that do not have a known genetic basis for resistance can exhibit profoundly reduced susceptibility when they form biofilms (48).It was reported previously that changes in gene expression induced a biofilm-specific phenotype (5, 13, 22, 35, 41, 42). Several genes have been proposed to be particularly important for biofilm formation, and the importance of the rpoS gene in Escherichia coli biofilm formation was suggested recently (1, 10, 22, 42). It has been suggested that induction of an rpoS-mediated stress response results in physiological changes that could contribute to antibiotic resistance (29). Although several mechanisms and genes have been proposed to explain biofilm resistance to antibiotics, this resistance is not still fully understood because these mechanisms seem to work together within a biofilm community. In addition, the physiology of biofilm cells is remarkably heterogeneous and varies according to the location of individual cells within biofilms (33, 34, 46).In this study, susceptibility of E. coli cells in biofilms to antibiotics was investigated. The E. coli cells in the deeper layers of mature biofilms were directly treated with three antibiotics with different molecular targets, the β-lactam ampicillin, the aminoglycoside kanamycin, and the fluoroquinolone ofloxacin. The biofilm biomass was removed before antibiotic treatment, and only the cells located in the deeper layers of the mature biofilms were directly exposed to antibiotics; thus, the effects of restricted antibiotic and nutrient penetration, as well as heterogeneous physiological states in biofilms, were reduced. Although ofloxacin and kanamycin effectively killed the biofilm cells, ampicillin could not kill the cells, which led to regrowth of biofilms. However, the cells in young colony biofilms were completely killed by ampicillin. Therefore, to determine which genes are induced in the mature biofilm cells, allowing increased resistance to ampicillin, global gene expression was analyzed at different stages of biofilm formation, the attachment, colony formation, and maturation stages. Based on the experimental data obtained, possible mechanisms of the increased biofilm resistance to ampicillin are discussed below.  相似文献   

2.
Biofilms are sessile microbial communities that cause serious chronic infections with high morbidity and mortality. In order to develop more effective approaches for biofilm control, a series of linear cationic antimicrobial peptides (AMPs) with various arginine (Arg or R) and tryptophan (Trp or W) repeats [(RW)n-NH2, where n = 2, 3, or 4] were rigorously compared to correlate their structures with antimicrobial activities affecting the planktonic growth and biofilm formation of Escherichia coli. The chain length of AMPs appears to be important for inhibition of bacterial planktonic growth, since the hexameric and octameric peptides significantly inhibited E. coli growth, while tetrameric peptide did not cause noticeable inhibition. In addition, all AMPs except the tetrameric peptide significantly reduced E. coli biofilm surface coverage and the viability of biofilm cells, when added at inoculation. In addition to inhibition of biofilm formation, significant killing of biofilm cells was observed after a 3-hour treatment of preformed biofilms with hexameric peptide. Interestingly, treatment with the octameric peptide caused significant biofilm dispersion without apparent killing of biofilm cells that remained on the surface; e.g., the surface coverage was reduced by 91.5 ± 3.5% by 200 μM octameric peptide. The detached biofilm cells, however, were effectively killed by this peptide. Overall, these results suggest that hexameric and octameric peptides are potent inhibitors of both bacterial planktonic growth and biofilm formation, while the octameric peptide can also disperse existing biofilms and kill the detached cells. These results are helpful for designing novel biofilm inhibitors and developing more effective therapeutic methods.Antimicrobial peptides (AMPs) are promising alternatives to traditional antibiotics (5). Native AMPs are part of the host defense in organisms ranging from bacteria to insects, plants, and animals (14). They are capable of eliminating a broad spectrum of microorganisms, including viruses, bacteria, and fungi (4, 14). Compared with widespread antibiotic resistance (38), resistance to AMPs is rare, possibly because AMPs directly target cell membranes that are essential to microbes (14, 29). In addition, no cross-resistance has been observed in clinic due to the diversity of peptide sequences (42). Thus, native and synthetic AMPs offer potential alternatives to antibiotics for treating drug-resistant infections (3, 26, 27).In mammalian innate immune systems, some AMPs are produced constitutively, while others are inducible within hours after detection of invading microorganisms (4, 13). Although the detailed mechanism of AMPs'' activities remains elusive (5), AMPs are known to disrupt cell membranes of microbes, interfere with metabolism, and/or target cytoplasmic components (41). Most known AMPs are cationic and amphiphilic (29). It is hypothesized that the initial interaction occurs via an electrostatic attraction between the AMP molecule and microbial membrane. Cationic AMPs can cover bacterial membranes, disrupt the membrane potential, create pores across the membrane, and consequently cause the leak of cell contents and cell death (27, 41). AMPs are relatively selective in targeting microbes rather than mammalian cells, most likely because of the fundamental differences between microbial and host membranes (41), e.g., a higher abundance of negatively charged phospholipids and an absence of cholesterol in microbial membranes.Known AMPs vary dramatically in sequence, size (from 12 to 50 amino acids), and structure (α-helices or β-sheets) (23). However, most AMPs have two types of side chains with relatively conservative sequences: positively charged basic residues, containing arginine (R), lysine (K), and/or histidine (H), that presumably mediate the interaction with the negatively charged microbial membrane, and bulky hydrophobic residues, rich in tryptophan (W), proline (P), and/or phenylalanine (F), that facilitate permeabilization and membrane disruption (26).Although AMPs are promising agents for antimicrobial therapies (15), only a few have made it to clinical trials and applications, with varied success (15, 42). There are several issues that need further development. First, the MICs of AMPs are relatively high compared to those of conventional antibiotics. Recent studies suggest that the peptide/lipid (P/L) ratio needs to be higher than a threshold to allow the AMPs to be oriented perpendicular to the membrane so that pores can be created to kill bacteria (22, 30). Thus, an optimization of peptide structure and size may improve their antimicrobial activities. In addition to the high MICs, the wide application of AMPs is also hindered by their high manufacturing costs and the cytotoxicity of some AMPs.Given the limit of currently available AMPs, it is important to develop more effective AMPs with reduced manufacturing cost and enhanced activity (17, 26, 28, 39). Strøm et al. (39) chemically synthesized a series of short cationic AMPs containing repeating R and W residues in order to identify the minimal pharmacophore with high antimicrobial activities. The data suggest that tetrapeptides or capped tripeptides are effective and there is no correlation between the order of amino acids and antimicrobial activity. Liu et al. (26) analyzed the effects of chain length on the activities of AMPs with repeating pharmacophore sequences (RW)n-NH2 (n = 1, 2, 3, 4, or 5). The tests of antimicrobial activities and the hemolysis of red blood cells suggest that (RW)3-NH2 has the optimal chain length. Although longer chains are more potent antimicrobials, they can stimulate hemolysis.Most of the AMP studies to date are focused on planktonic bacteria. However, the majority of pathogenic bacteria tend to adhere to surfaces and form sessile microbial communities with highly hydrated structures of secreted polysaccharide matrix, collectively known as biofilms (9). Biofilms can tolerate up to 1,000 times more antibiotics and disinfectants than their planktonic counterparts (2, 7, 8). For example, Folkesson et al. (12) reported that biofilm formation of E. coli K-12 increases its tolerance to polymyxin E, a polypeptide antibiotic that kills Gram-negative bacteria by disrupting membranes (34, 40). Since biofilms are involved in 80% of human bacterial infections (1), it is necessary to study biofilm inhibition and dispersion by AMPs.In this study, a series of linear peptides (RW)n-NH2 (where n = 2, 3, or 4) were studied for the effects of their activities on planktonic cells and biofilms of E. coli to understand the structural effects on the antimicrobial activities of AMPs. We chose E. coli RP437 in this study because it is one of the model strains for biofilm research and allows us to compare the data with those of our previous studies (6, 16, 19, 20).  相似文献   

3.
The stochastic Ricker population model was used to investigate the generation and maintenance of genetic diversity in a bacterial population grown in a spatially structured environment. In particular, we showed that Escherichia coli undergoes dramatic genetic diversification when grown as a biofilm. Using a novel biofilm entrapment method, we retrieved 64 clones from each of six different depths of a mature biofilm, and after subculturing for ∼30 generations, we measured their growth kinetics in three different media. We fit a stochastic Ricker population growth model to the recorded growth curves. The growth kinetics of clonal lineages descendant from cells sampled at different biofilm depths varied as a function of both the depth in the biofilm and the growth medium used. We concluded that differences in the growth dynamics of clones were heritable and arose during adaptive evolution under local conditions in a spatially heterogeneous environment. We postulate that under nutrient-limited conditions, selective sweeps would be protracted and would be insufficient to purge less-fit variants, a phenomenon that would allow the coexistence of genetically distinct clones. These findings contribute to the current understanding of biofilm ecology and complement current hypotheses for the maintenance and generation of microbial diversity in spatially structured environments.The mechanisms that lead to the genesis and maintenance of diversity in communities have intrigued geneticists and ecologists alike for decades (6, 17, 27, 33, 39, 49). This is particularly challenging for microbial communities, in which ecological and evolutionary processes occur on roughly the same time scale (3, 16, 38) and where the outcome of these processes may be affected by the spatial structure in which these communities grow.Bacterial biofilms are examples of spatially structured communities that have been the subject of intense research in medical and engineering contexts in recent years (3, 8, 20, 48, 56). Previous work has shown that the phenotypic characteristics of bacterial populations in biofilms are distinct from those of their free-swimming counterparts (8). These bacterial assemblages form physically and chemically heterogeneous structures (20) whose complex architecture strongly influences mass transfer (56). This results in the formation of steep gradients of nutrients, waste products, pH, redox potential, and electron acceptors, which results in the creation of distinct and perhaps unique niches on a microscale. This places selective pressure on variants that have enhanced fitness and are well adapted to local conditions. From a theoretical perspective, this would be expected to increase genetic diversity within a population by precluding competitive exclusion, yet this has not previously been demonstrated empirically.The degree of diversification that occurs within populations growing in biofilms is not well understood, nor are the spatial and temporal dynamics of bacterial species succession in biofilms. However, it is known that the physical and chemical heterogeneity of microbial biofilms has profound effects on microbial growth and activity. Most bacterial cells in biofilms are not highly active and grow slowly if at all. For example, active protein synthesis occurs only in the uppermost zone (32 ± 3 μm) of Pseudomonas aeruginosa biofilms (4). Likewise, in Klebsiella pneumoniae biofilms, fast growth occurs near the interface of the biofilm and bulk fluid, and cells inside the biofilm show little growth (55). The near absence of growth in interior regions of biofilms may lead to an increased tempo of diversification, since numerous studies have shown that mutation frequencies are elevated in slowly growing cells (28). If this occurs within a biofilm, then clones might exhibit a high genotypic variability that could have significant practical implications in terms of yielding spontaneous mutants that are resistant to antimicrobial agents.Experimental evolution has contributed greatly to our understanding of the causes and consequences of genetic diversity in populations (reviewed in references 23, 29, and 42). Initially, research focused on characterizing diversity within populations that evolved in spatially homogenous environments (e.g., chemostat and batch systems) (13, 15, 19, 30-32, 45, 47, 50-53). Several studies have highlighted a role for spatial heterogeneity in the emergence and maintenance of genetic diversity (25, 26, 43). Korona and colleagues (25, 26) compared populations that evolved in batch cultures to populations that evolved with a spatial structure and demonstrated that phenotypic diversity was greatest with spatial structure. In other work, Rainey and Travisano (43) showed that populations of Pseudomonas grown in static broth microcosms diversified so that some ecotypes occupied a floating biofilm on the surface of the broth while others occupied the liquid phase or glass surface of the culture. Boles et al. (2, 3) investigated the extent of diversification of Pseudomonas using biofilms that evolved in flow-cell systems. They reported that genetic changes produced by a recA-dependent mechanism affected multiple traits, with some biofilm-derived variants being better able to disseminate while others were better able to form biofilms (3). Further study showed that in some cells, endogenous oxidative stress caused double-stranded DNA breaks that when repaired by recombinatorial DNA repair genes gave rise to mutations (2). These previous studies demonstrate the pivotal role of spatial structure in the generation and maintenance of diversity in evolving bacterial populations.In this study, we extended this work by using novel techniques to characterize diversity in Escherichia coli biofilms that allowed us to recover clones from specific depths within a biofilm. The growth kinetics of clones from six different biofilm depths were measured and modeled using an analysis-of-variance formulation of the stochastic Ricker model of population dynamics with environmental noise (11, 40). Rigorous statistical methods were used to show that after 1 month of cultivation, the extant diversity in E. coli biofilms was extraordinarily high and varied with depth.  相似文献   

4.
Molecules become readily visible by magnetic resonance imaging (MRI) when labeled with a paramagnetic tag. Consequently, MRI can be used to image their transport through porous media. In this study, we demonstrated that this method could be applied to image mass transport processes in biofilms. The transport of a complex of gadolinium and diethylenetriamine pentaacetic acid (Gd-DTPA), a commercially available paramagnetic molecule, was imaged both in agar (as a homogeneous test system) and in a phototrophic biofilm. The images collected were T1 weighted, where T1 is an MRI property of the biofilm and is dependent on Gd-DTPA concentration. A calibration protocol was applied to convert T1 parameter maps into concentration maps, thus revealing the spatially resolved concentrations of this tracer at different time intervals. Comparing the data obtained from the agar experiment with data from a one-dimensional diffusion model revealed that transport of Gd-DTPA in agar was purely via diffusion, with a diffusion coefficient of 7.2 × 10−10 m2 s−1. In contrast, comparison of data from the phototrophic biofilm experiment with data from a two-dimensional diffusion model revealed that transport of Gd-DTPA inside the biofilm was by both diffusion and advection, equivalent to a diffusion coefficient of 1.04 × 10−9 m2 s−1. This technology can be used to further explore mass transport processes in biofilms, either by using the wide range of commercially available paramagnetically tagged molecules and nanoparticles or by using bespoke tagged molecules.Biofilms are utilized in a wide range of biotechnological processes, such as cleansing municipal and industrial wastewater, bioremediation of hazardous waste sites, biofuel production, and the generation of electricity in microbial fuel cells (20, 31, 35). They also play an important role in mediating the geochemistry of the natural environment (35). Critically, our growing understanding of the biology, physics, and chemistry of biofilms is allowing us to manipulate biofilms and enhance their performance in a variety of biotechnologies (33). The optimization of biofilm processes is, however, hindered when a lack of quantitative measurements of critical biofilm parameters exists.For the biofilm to function, the relevant substrates must be transported through the biofilm matrix, where they are metabolized. The rate at which these metabolites are transported through the biofilm can be critical in controlling the performance of the biofilm (5, 8, 13, 31). Equally, the rate at which the biofilm can sequester nonmetabolizable pollutants, such as nonmetabolizable heavy metals and recalcitrant organics, is also mediated by the transport rate (9, 28). Previous studies of mass transport inside biofilms show that transport occurs not only by diffusion but also by advection if the biofilm contains interconnected channels (5, 9, 13, 19, 39, 40, 45). When transported by diffusion, the mass of the diffusing solute plays a key role in mediating the transport rate. That is, the higher the molecular mass of the solute, the lower its diffusion coefficient (7, 39). Moreover, the molecular masses and diffusion rates of these solutes vary considerably, ranging from low-mass, fast-diffusing metabolites, such as H2 and O2, to large, slowly diffusing organic macromolecules tens to hundreds of kDa in size. Indeed, high-molecular-mass molecules and nanoparticles are an important part of the substrate and pollutant load in both wastewater treatment and natural aquatic systems (21). At a certain size, large macromolecules and nanoparticles become too large to diffuse into the dense extracellular polymeric substance (EPS) matrix, although they still can be transported deep into the biofilm along open channels (9, 39).Moreover, due to the heterogeneous nature of biofilms, substrates can also display significant spatial variation in mass transport rates, such as a decrease in transport rate with biofilm depth (4). As attempts to understand biofilm function or enhance biofilm performance are dependent upon accurate mass transport data sets, quantifying the transport behaviors of different-molecular-mass molecules in different biofilms is key to allowing us to model real biofilm systems more accurately.Recognizing the importance of mass transport, researchers have already used a variety of methods, such as microelectrodes, confocal laser scanning microscopy (CLSM), fluorescence recovery after photobleaching (FRAP), and two-photon excitation microscopy to obtain mass transport data from biofilms (7, 11, 12). These approaches have provided invaluable data on mass transport within biofilms. However, as with any method, each has certain limitations. For example, microelectrodes are used to measure the mass transport of low-molecular-mass molecules; particulates and high-molecular-mass molecules are undetectable by this method. Moreover, the insertion of a probe is invasive and thus has potential to disrupt the surrounding material, altering results. This could be problematic when numerous insertions must be made, such as during spatial mapping of diffusion coefficients in heterogeneous biofilms. Conversely, CLSM is noninvasive. However, small molecules such as H2 or O2 cannot be labeled with the fluorescent probe, and thus only the transport of higher-molecular-weight compounds can be determined. This method, which relies on photons penetrating the biofilm, is limited both to biofilm thickness (<100 μm) and to its density due to optical scattering effects (26, 43). Although the two-photon excitation method can overcome the depth penetration limitation of CLSM by approximately four times (26), it is not suitable where biofilms exceed these thicknesses. FRAP also suffers similar thickness limitations and light-scattering effects. However, the capacity of magnetic resonance imaging (MRI) for completely noninvasive measurement of the transport of both low- and high-molecular-mass compounds and its ability to image inside hydrated biological matrices (1, 30), no matter what thickness, means that it has significant potential for mass transport analysis of biofilms and can thus be an invaluable additional tool in this research field.Researchers have already used MRI to examine flow dynamics over biofilm surfaces (22, 37), metabolite consumption and production (23), the flux of heavy metals in metal-immobilizing bioreactors (15, 25), water diffusion in biofilms (28, 44), and the transport and fate of metals both in natural and artificial biofilms (28, 29) and in real methanogenic granules which are employed in anaerobic wastewater treatment (2).  相似文献   

5.
6.
7.
Most microbes, including the fungal pathogen Cryptococcus neoformans, can grow as biofilms. Biofilms confer upon microbes a range of characteristics, including an ability to colonize materials such as shunts and catheters and increased resistance to antibiotics. Here, we provide evidence that coating surfaces with a monoclonal antibody to glucuronoxylomannan, the major component of the fungal capsular polysaccharide, immobilizes cryptococcal cells to a surface support and, subsequently, promotes biofilm formation. We used time-lapse microscopy to visualize the growth of cryptococcal biofilms, generating the first movies of fungal biofilm growth. We show that when fungal cells are immobilized using surface-attached specific antibody to the capsule, the initial stages of biofilm formation are significantly faster than those on surfaces with no antibody coating or surfaces coated with unspecific monoclonal antibody. Time-lapse microscopy revealed that biofilm growth was a dynamic process in which cells shuffled position during budding and was accompanied by emergence of planktonic variant cells that left the attached biofilm community. The planktonic variant cells exhibited mobility, presumably by Brownian motion. Our results indicate that microbial immobilization by antibody capture hastens biofilm formation and suggest that antibody coating of medical devices with immunoglobulins must exclude binding to common pathogenic microbes and the possibility that this effect could be exploited in industrial microbiology.Cryptococcus neoformans is a fungal pathogen that is ubiquitous in the environment and enters the body via the inhalation of airborne particles. The C. neoformans cell is surrounded by a layer of polysaccharide that consists predominantly of glucuronoxylomannan (GXM), which forms a protective capsule around the microbe. The capsule has been shown to be essential for virulence in murine models of infection (5-7) and, thus, is considered a key virulence factor. C. neoformans is the causative agent of cryptococcosis, a disease that primarily affects individuals with impaired immune systems, and is a significant problem in AIDS patients (21, 31). The most common manifestation of cryptococcosis is meningoencephalitis.Biofilms are communities of microbes that are attached to surfaces and held together by an extracellular matrix, often consisting predominantly of polysaccharides (8, 10). A great deal is known about bacterial biofilms (3, 9, 24, 30), but fungal biofilm formation is much less studied. Candida albicans is known to synthesize biofilms (11, 28, 29), as is C. neoformans. Biofilm-like structures consisting of innumerable cryptococcal cells encased in a polysaccharide matrix have been reported in human cases of cryptococcosis (32). Biofilm formation confers upon the microbe the capacity for drug resistance, and microbial cells in biofilms are less susceptible to host defense mechanisms (2, 4, 9, 12). In this regard, cells within C. neoformans biofilms are significantly less susceptible to caspofungin and amphotericin B than are planktonic cells (19). The cells within the biofilm are also resistant to the actions of fluconazole and voriconazole and various microbial oxidants and peptides (17, 19).Bacterial and fungal biofilms form readily on prosthetic materials, which poses a tremendous risk of chronic infection (10, 13, 15, 27). C. neoformans biofilms can form on a range of surfaces, including glass, polystyrene, and polyvinyl, and material devices, such as catheters (16). C. neoformans can form biofilms on the ventriculoatrial shunts used to decompress intracerebral pressure in patients with cryptococcal meningoencephalitis (32).The polysaccharide capsule of C. neoformans is essential for biofilm formation (18), and biofilm formation involves the shedding and accumulation of large amounts of GXM into the biofilm extracellular matrix (16). Previously, we reported that antibody to GXM in solution could inhibit biofilm formation through a process that presumably involves interference with polysaccharide shedding (18, 20). However, the effect of antibody-mediated immobilization of C. neoformans cells on cryptococcal biofilm formation has not been explored. In this paper, we report that the monoclonal antibody (MAb) 18B7, which is specific for the capsular polysaccharide GXM, can capture and immobilize C. neoformans to surfaces, a process that promotes biofilm formation. Interestingly, we identified planktonic variant C. neoformans cells that appeared to escape from the biofilm, but whose functions are not known. The results provide new insights on biofilm formation.  相似文献   

8.
9.
10.
Iron is an essential element for life but also serves as an environmental signal for biofilm development in the opportunistic human pathogen Pseudomonas aeruginosa. Under iron-limiting conditions, P. aeruginosa displays enhanced twitching motility and forms flat unstructured biofilms. In this study, we present evidence suggesting that iron-regulated production of the biosurfactant rhamnolipid is important to facilitate the formation of flat unstructured biofilms. We show that under iron limitation the timing of rhamnolipid expression is shifted to the initial stages of biofilm formation (versus later in biofilm development under iron-replete conditions) and results in increased bacterial surface motility. In support of this observation, an rhlAB mutant defective in biosurfactant production showed less surface motility under iron-restricted conditions and developed structured biofilms similar to those developed by the wild type under iron-replete conditions. These results highlight the importance of biosurfactant production in determining the mature structure of P. aeruginosa biofilms under iron-limiting conditions.The biofilm mode of bacterial growth is a surface-attached state in which cells are closely packed and encased in an extracellular polymeric matrix (10, 27). Biofilms are abundant in nature and are of clinical, environmental, and industrial importance. Biofilm development is known to follow a series of complex but discrete and tightly regulated steps (18, 27), including (i) microbial attachment to the surface, (ii) growth and aggregation of cells into microcolonies, (iii) maturation, and (iv) dissemination of progeny cells that can colonize new niches. Over the last decade, several key processes important for biofilm formation have been identified, including quorum sensing (12) and surface motility (28).One of the best-studied model organisms for biofilm development is the bacterium Pseudomonas aeruginosa (10), a notorious opportunistic pathogen which causes many types of infections, including biofilm-associated chronic lung infections in individuals with cystic fibrosis (10, 24, 41). Like most organisms, P. aeruginosa requires iron for growth, as iron serves as a cofactor for enzymes that are involved in many basic cellular functions and metabolic pathways. Recent work has shown that at iron concentrations that are not limiting for growth, this metal serves as a signal for biofilm development (40). Iron limitation imposed, for example, by the mammalian iron chelator lactoferrin blocks the ability of P. aeruginosa biofilms to mature from thin layers of cells attached to a surface into large multicellular mushroom-like biofilm structures (40). By chelating iron, lactoferrin induces twitching motility (a specialized form of surface motility), which causes the cells to move across the surface instead of settling down to form structured communities (39, 40). In a recent paper, Berlutti et al. (5) provided further support for the role of iron in cell aggregation and biofilm formation. They reported that in the liquid phase, iron limitation induced motility and transition to the free-living (i.e., planktonic) mode of growth, while increased iron concentrations facilitated cell aggregation and biofilm formation. We recently demonstrated that iron limitation-induced twitching motility is regulated by quorum sensing (31). Quorum sensing allows bacteria to sense and respond to their population density via the production of small diffusible signal molecules. In P. aeruginosa and many other Gram-negative bacteria, these signal molecules are N-acyl homoserine lactones (acyl-HSLs), which have specific receptors (R proteins) (16, 30). P. aeruginosa possesses two acyl-HSL quorum-sensing systems, one for production of and response to N-3-oxo-dodecanoyl homoserine lactone (3OC12-HSL) (LasR-LasI) and the other for production of and response to N-butanoyl homoserine lactone (C4-HSL) (RhlR-RhlI) (35, 37). We have reported that an rhlI mutant unable to synthesize the C4-HSL signal was impaired in iron limitation-induced twitching motility and formed structured biofilms under iron-limiting conditions (31).The correlation between twitching motility, the RhlR-RhlI quorum-sensing system, and iron-regulated biofilm formation led us to hypothesize that rhamnolipids are involved in mediating this process. Rhamnolipids are surface-active amphipathic molecules composed of a hydrophobic lipid and a hydrophilic sugar moiety and compose the main constituents of the biosurfactant produced by P. aeruginosa (reviewed in reference 42). The biosurfactant is required for a form of surface motility called swarming, where it functions as a wetting agent and reduces surface tension (8, 14). Furthermore, elements constituting the biosurfactant were recently shown to modulate the swarming behavior by acting as chemotactic-like stimuli (43). Rhamnolipids are also important in maintaining biofilm structure and inducing biofilm dispersion (6, 11, 29). Their synthesis requires the expression of the rhlAB operon, which is regulated by the RhlR-RhlI quorum-sensing system (14, 25, 32) and is also induced under iron-limiting conditions (14).In this study, we test this hypothesis and demonstrate that rhamnolipid production is induced under iron-limiting conditions and that this promotes twitching motility. We found that increased expression of rhamnolipid synthesis genes during early biofilm development under iron-limiting conditions induces surface motility and results in formation of a thin flat biofilm. Furthermore, a mutant that is incapable of synthesizing rhamnolipids does not display twitching motility under iron-limiting conditions and thus forms structured biofilms under these conditions. These results highlight the importance of biosurfactant production in determining the architecture of mature P. aeruginosa biofilms under iron-limiting conditions.  相似文献   

11.
12.
Human cytomegalovirus (HCMV) UL37 proteins traffic sequentially from the endoplasmic reticulum (ER) to the mitochondria. In transiently transfected cells, UL37 proteins traffic into the mitochondrion-associated membranes (MAM), the site of contact between the ER and mitochondria. In HCMV-infected cells, the predominant UL37 exon 1 protein, pUL37x1, trafficked into the ER, the MAM, and the mitochondria. Surprisingly, a component of the MAM calcium signaling junction complex, cytosolic Grp75, was increasingly enriched in heavy MAM from HCMV-infected cells. These studies show the first documented case of a herpesvirus protein, HCMV pUL37x1, trafficking into the MAM during permissive infection and HCMV-induced alteration of the MAM protein composition.The human cytomegalovirus (HCMV) UL37 immediate early (IE) locus expresses multiple products, including the predominant UL37 exon 1 protein, pUL37x1, also known as viral mitochondrion-localized inhibitor of apoptosis (vMIA), during lytic infection (16, 22, 24, 39, 44). The UL37 glycoprotein (gpUL37) shares UL37x1 sequences and is internally cleaved, generating pUL37NH2 and gpUL37COOH (2, 22, 25, 26). pUL37x1 is essential for the growth of HCMV in humans (17) and for the growth of primary HCMV strains (20) and strain AD169 (14, 35, 39, 49) but not strain TownevarATCC in permissive human fibroblasts (HFFs) (27).pUL37x1 induces calcium (Ca2+) efflux from the endoplasmic reticulum (ER) (39), regulates viral early gene expression (5, 10), disrupts F-actin (34, 39), recruits and inactivates Bax at the mitochondrial outer membrane (MOM) (4, 31-33), and inhibits mitochondrial serine protease at late times of infection (28).Intriguingly, HCMV UL37 proteins localize dually in the ER and in the mitochondria (2, 9, 16, 17, 24-26). In contrast to other characterized, similarly localized proteins (3, 6, 11, 23, 30, 38), dual-trafficking UL37 proteins are noncompetitive and sequential, as an uncleaved gpUL37 mutant protein is ER translocated, N-glycosylated, and then imported into the mitochondria (24, 26).Ninety-nine percent of ∼1,000 mitochondrial proteins are synthesized in the cytosol and directly imported into the mitochondria (13). However, the mitochondrial import of ER-synthesized proteins is poorly understood. One potential pathway is the use of the mitochondrion-associated membrane (MAM) as a transfer waypoint. The MAM is a specialized ER subdomain enriched in lipid-synthetic enzymes, lipid-associated proteins, such as sigma-1 receptor, and chaperones (18, 45). The MAM, the site of contact between the ER and the mitochondria, permits the translocation of membrane-bound lipids, including ceramide, between the two organelles (40). The MAM also provides enriched Ca2+ microdomains for mitochondrial signaling (15, 36, 37, 43, 48). One macromolecular MAM complex involved in efficient ER-to-mitochondrion Ca2+ transfer is comprised of ER-bound inositol 1,4,5-triphosphate receptor 3 (IP3R3), cytosolic Grp75, and a MOM-localized voltage-dependent anion channel (VDAC) (42). Another MAM-stabilizing protein complex utilizes mitofusin 2 (Mfn2) to tether ER and mitochondrial organelles together (12).HCMV UL37 proteins traffic into the MAM of transiently transfected HFFs and HeLa cells, directed by their NH2-terminal leaders (8, 47). To determine whether the MAM is targeted by UL37 proteins during infection, we fractionated HCMV-infected cells and examined pUL37x1 trafficking in microsomes, mitochondria, and the MAM throughout all temporal phases of infection. Because MAM domains physically bridge two organelles, multiple markers were employed to verify the purity and identity of the fractions (7, 8, 19, 46, 47).(These studies were performed in part by Chad Williamson in partial fulfillment of his doctoral studies in the Biochemistry and Molecular Genetics Program at George Washington Institute of Biomedical Sciences.)HFFs and life-extended (LE)-HFFs were grown and not infected or infected with HCMV (strain AD169) at a multiplicity of 3 PFU/cell as previously described (8, 26, 47). Heavy (6,300 × g) and light (100,000 × g) MAM fractions, mitochondria, and microsomes were isolated at various times of infection and quantified as described previously (7, 8, 47). Ten- or 20-μg amounts of total lysate or of subcellular fractions were resolved by SDS-PAGE in 4 to 12% Bis-Tris NuPage gels (Invitrogen) and examined by Western analyses (7, 8, 26). Twenty-microgram amounts of the fractions were not treated or treated with proteinase K (3 μg) for 20 min on ice, resolved by SDS-PAGE, and probed by Western analysis. The blots were probed with rabbit anti-UL37x1 antiserum (DC35), goat anti-dolichyl phosphate mannose synthase 1 (DPM1), goat anti-COX2 (both from Santa Cruz Biotechnology), mouse anti-Grp75 (StressGen Biotechnologies), and the corresponding horseradish peroxidase-conjugated secondary antibodies (8, 47). Reactive proteins were detected by enhanced chemiluminescence (ECL) reagents (Pierce), and images were digitized as described previously (26, 47).  相似文献   

13.
14.
Among the most difficult bacterial infections encountered in treating patients are wound infections, which may occur in burn victims, patients with traumatic wounds, necrotic lesions in people with diabetes, and patients with surgical wounds. Within a wound, infecting bacteria frequently develop biofilms. Many current wound dressings are impregnated with antimicrobial agents, such as silver or antibiotics. Diffusion of the agent(s) from the dressing may damage or destroy nearby healthy tissue as well as compromise the effectiveness of the dressing. In contrast, the antimicrobial agent selenium can be covalently attached to the surfaces of a dressing, prolonging its effectiveness. We examined the effectiveness of an organoselenium coating on cellulose discs in inhibiting Pseudomonas aeruginosa and Staphylococcus aureus biofilm formation. Colony biofilm assays revealed that cellulose discs coated with organoselenium completely inhibited P. aeruginosa and S. aureus biofilm formation. Scanning electron microscopy of the cellulose discs confirmed these results. Additionally, the coating on the cellulose discs was stable and effective after a week of incubation in phosphate-buffered saline. These results demonstrate that 0.2% selenium in a coating on cellulose discs effectively inhibits bacterial attachment and biofilm formation and that, unlike other antimicrobial agents, longer periods of exposure to an aqueous environment do not compromise the effectiveness of the coating.Among the most difficult bacterial infections encountered in treating patients are wound infections, which may occur in burn victims (10), patients with traumatic wounds (33), people with diabetes (27), and patients with surgical wounds (29, 31). Two of the more common causative agents of wound infections are Staphylococcus aureus and Pseudomonas aeruginosa (10, 27, 29, 31, 33). Such infections often lead to fatality; the mortality rate among patients infected with P. aeruginosa ranges from 26% to 55% (9, 49), while mortality from S. aureus infection ranges from 19% to 38% (28, 46, 50). As opportunistic pathogens, S. aureus and P. aeruginosa cause few infections in healthy individuals but readily cause infection once host defenses are compromised, such as with the removal of skin from burns (10). S. aureus infection originates from the normal flora of either the patient or health care workers (48), while P. aeruginosa is acquired from the environment surrounding the patient (41). Once established on the skin, S. aureus and P. aeruginosa are then able to colonize the wound. Infection results if the organisms proliferate in the wound environment.Both P. aeruginosa and S. aureus often exist within burn wounds as biofilms (43, 47). A biofilm is presently defined as a sessile microbial community characterized by cells that are irreversibly attached either to a substratum or to each other (16). Biofilms, which can attain over 100 μm in thickness, are made up of multiple layers of bacteria in an exopolysaccharide matrix (12, 16, 42). Sauer et al. showed that P. aeruginosa biofilms form in distinct developmental stages: reversible attachment, irreversible attachment, two stages of maturation, and a dispersion phase (42). Clinically, biofilms present serious medical management problems through their association with different chronic infections (37). During vascular catheter-related infections and sepsis, biofilms serve as a reservoir of bacteria from which planktonic cells detach and spread throughout the tissue and/or enter the circulatory system, resulting in bacteremia or septicemia (15). Factors specific to the bacterium may influence the formation of bacterial biofilms at different infection sites or surfaces. For example, during the initial attachment stage the flagellum, lipopolysaccharide, and possibly outer membrane proteins play a major role in bringing P. aeruginosa into proximity with the surface as well as mediating the interaction with the substratum (12). Using the murine model of thermal injury, we recently showed that P. aeruginosa forms a biofilm within the thermally injured tissues (43). Clinically, the surgeons debride the infected or dead tissues; however, a few microorganisms may remain on the tissue surface and reinitiate biofilm formation.Antibiotics, silver, or chitosan, attached to or embedded in gauze, have been shown to be efficacious in preventing wound infection (21, 24, 26, 36). However, the resistance of P. aeruginosa and S. aureus to available antibiotics severely limits the choices for antibiotic treatment (13, 40). Additionally, silver compounds, such as silver nitrate and silver sulfadiazine, leaching from dressings are toxic to human fibroblasts even at low concentrations (20, 25). Thus, effective alternative antimicrobial agents that contact the thermally injured/infected tissues and prevent the development of bacterial biofilms are required. Previous studies have shown that selenium (Se) can be covalently bound to a solid matrix and retain its ability to catalyze the formation of superoxide radicals (O2·−) (30). These superoxide radicals inhibit bacterial attachment to the solid surface (30). In this study, we examined the ability of a newly synthesized organoselenium-methacrylate polymer (Se-MAP) to block biofilm formation by both S. aureus and P. aeruginosa. These bacteria were chosen since they cause a major share of wound infections and because drug-resistant forms of these bacteria have become a serious problem in the treatment and management of these wound infections (6, 13, 17, 18, 38). Results of the study show that 0.2% (wt/wt) Se in Se-MAP covalently attached to cellulose discs inhibited P. aeruginosa and S. aureus biofilm formation. This could lead to the development of a selenium-based antimicrobial coating for cotton materials that will prevent the bacterial attachment and colonization that can ultimately lead to bacterial biofilm formation during chronic infections.  相似文献   

15.
16.
Extracellular DNA (eDNA) is produced by several bacterial species and appears to contribute to biofilm development and cell-cell adhesion. We present data showing that the oral commensals Streptococcus sanguinis and Streptococcus gordonii release DNA in a process induced by pyruvate oxidase-dependent production of hydrogen peroxide (H2O2). Surprisingly, S. sanguinis and S. gordonii cell integrity appears unaffected by conditions that cause autolysis in other eDNA-producing bacteria. Exogenous H2O2 causes release of DNA from S. sanguinis and S. gordonii but does not result in obvious lysis of cells. Under DNA-releasing conditions, cell walls appear functionally intact and ribosomes are retained over time. During DNA release, intracellular RNA and ATP are not coreleased. Hence, the release mechanism appears to be highly specific for DNA. Release of DNA without detectable autolysis is suggested to be an adaptation to the competitive oral biofilm environment, where autolysis could create open spaces for competitors to invade. Since eDNA promotes cell-to-cell adhesion, release appears to support oral biofilm formation and facilitates exchange of genetic material among competent strains.The release of bacterial DNA into the environment is of recent interest since this polymer is now recognized to stabilize cell-to-cell adherence and biofilm architecture (1, 35, 37). Treatment of extracellular DNA (eDNA) with DNase results in reduced intercellular stickiness, consistent with an adhesive function for eDNA. Furthermore, eDNA from Neisseria meningitis appears to have sufficient structural integrity to transform competent strains (11), indicating chromosomal origin. Since the abundance of eDNA is influenced by growth conditions, DNA release can also be regulated (40).DNA release is typically a consequence of cell lysis. Linked to DNA release, genetic transformation is the natural ability of competent bacterial species to take up DNA from the environment (13, 34, 42). During competence development, Streptococcus pneumoniae DNA is released by lysis of a subpopulation of cells (30, 42). Cell lysis and DNA release are controlled in a cell density-dependent signal transduction process. The S. pneumoniae comX regulon, carrying late competence genes, also includes the murein hydrolase genes lytA and cbpD (19, 42). Murein hydrolases digest structural components of the peptidoglycan, contributing to remodeling, recycling, and daughter cell separation. Furthermore, murein hydrolases trigger autolytic cell wall digestion, leading to release of DNA and other cellular content into the environment (36). The autolysis of bacterial cells as part of a regulated death program seems to be an important source for eDNA in diverse species, including Staphylococcus aureus (4, 36, 37), Staphylococcus epidermidis (35), Enterococcus faecalis (44), and Pseudomonas aeruginosa (1). In these species, the eDNA contributes to biofilm formation as a component of the extracellular biofilm matrix (35, 37, 44).Unlike for cell lysis-dependent release, the oral streptococci appear to induce eDNA release by a novel mechanism. In dual-species cultures, the oral commensals Streptococcus sanguinis and Streptococcus gordonii release eDNA in a manner dependent on pyruvate oxidase (Pox) generation of hydrogen peroxide (H2O2) under the control of ambient oxygen (23). In this report, we now provide direct evidence of selective H2O2-induced eDNA release by these oral commensal streptococci.  相似文献   

17.
Claudin-1, a component of tight junctions between liver hepatocytes, is a hepatitis C virus (HCV) late-stage entry cofactor. To investigate the structural and functional roles of various claudin-1 domains in HCV entry, we applied a mutagenesis strategy. Putative functional intracellular claudin-1 domains were not important. However, we identified seven novel residues in the first extracellular loop that are critical for entry of HCV isolates drawn from six different subtypes. Most of the critical residues belong to the highly conserved claudin motif W30-GLW51-C54-C64. Alanine substitutions of these residues did not impair claudin-1 cell surface expression or lateral protein interactions within the plasma membrane, including claudin-1-claudin-1 and claudin-1-CD81 interactions. However, these mutants no longer localized to cell-cell contacts. Based on our observations, we propose that cell-cell contacts formed by claudin-1 may generate specialized membrane domains that are amenable to HCV entry.Hepatitis C virus (HCV) is a major human pathogen that affects approximately 3% of the global population, leading to cirrhosis and hepatocellular carcinoma in chronically infected individuals (5, 23, 42). Hepatocytes are the major target cells of HCV (11), and entry follows a complex cascade of interactions with several cellular factors (6, 8, 12, 17). Infectious viral particles are associated with lipoproteins and initially attach to target cells via glycosaminoglycans and the low-density lipoprotein receptor (1, 7, 31). These interactions are followed by direct binding of the E2 envelope glycoprotein to the scavenger receptor class B type I (SR-B1) and then to the CD81 tetraspanin (14, 15, 33, 36). Early studies showed that CD81 and SR-B1 were necessary but not sufficient for HCV entry, and claudin-1 was discovered to be a requisite HCV entry cofactor that appears to act at a very late stage of the process (18).Claudin-1 is a member of the claudin protein family that participates in the formation of tight junctions between adjacent cells (25, 30, 37). Tight junctions regulate the paracellular transport of solutes, water, and ions and also generate apical-basal cell polarity (25, 37). In the liver, the apical surfaces of hepatocytes form bile canaliculi, whereas the basolateral surfaces face the underside of the endothelial layer that lines liver sinusoids. Claudin-1 is highly expressed in tight junctions formed by liver hepatocytes as well as on all hepatoma cell lines that are permissive to HCV entry (18, 24, 28). Importantly, nonhepatic cell lines that are engineered to express claudin-1 become permissive to HCV entry (18). Claudin-6 and -9 are two other members of the human claudin family that enable HCV entry into nonpermissive cells (28, 43).The precise role of claudin-1 in HCV entry remains to be determined. A direct interaction between claudins and HCV particles or soluble E2 envelope glycoprotein has not been demonstrated (18; T. Dragic, unpublished data). It is possible that claudin-1 interacts with HCV entry receptors SR-B1 or CD81, thereby modulating their ability to bind to E2. Alternatively, claudin-1 may ferry the receptor-virus complex to fusion-permissive intracellular compartments. Recent studies show that claudin-1 colocalizes with the CD81 tetraspanin at the cell surface of permissive cell lines (22, 34, 41). With respect to nonpermissive cells, one group observed that claudin-1 was predominantly intracellular (41), whereas another reported associations of claudin-1 and CD81 at the cell surface, similar to what is observed in permissive cells (22).Claudins comprise four transmembrane domains along with two extracellular loops and two cytoplasmic domains (19, 20, 25, 30, 37). The first extracellular loop (ECL1) participates in pore formation and influences paracellular charge selectivity (25, 37). It has been shown that the ECL1 of claudin-1 is required for HCV entry (18). All human claudins comprise a highly conserved motif, W30-GLW51-C54-C64, in the crown of ECL1 (25, 37). The exact function of this domain is unknown, and we hypothesized that it is important for HCV entry. The second extracellular loop is required for the holding function and oligomerization of the protein (25). Claudin-1 also comprises various signaling domains and a PDZ binding motif in the intracellular C terminus that binds ZO-1, another major component of tight junctions (30, 32, 37). We further hypothesized that some of these domains may play a role in HCV entry.To understand the role of claudin-1 in HCV infection, we developed a mutagenesis strategy targeting the putative sites for internalization, glycosylation, palmitoylation, and phosphorylation. The functionality of these domains has been described by others (4, 16, 25, 35, 37, 40). We also mutagenized charged and bulky residues in ECL1, including all six residues within the highly conserved motif W30-GLW51-C54-C64. None of the intracellular domains were found to affect HCV entry. However, we identified seven residues in ECL1 that are critical for entry mediated by envelope glycoproteins derived from several HCV subtypes, including all six residues of the conserved motif. These mutants were still expressed at the cell surface and able to form lateral homophilic interactions within the plasma membrane as well as to engage in lateral interactions with CD81. In contrast, they no longer engaged in homophilic trans interactions at cell-cell contacts. We conclude that the highly conserved motif W30-GLW51-C54-C64 of claudin-1 is important for HCV entry into target cells and participates in the formation of cell-cell contacts.  相似文献   

18.
Control of biofilms requires rapid methods to identify compounds effective against them and to isolate resistance-compromised mutants for identifying genes involved in enhanced biofilm resistance. While rapid screening methods for microtiter plate well (“static”) biofilms are available, there are no methods for such screening of continuous flow biofilms (“flow biofilms”). Since the latter biofilms more closely approximate natural biofilms, development of a high-throughput (HTP) method for screening them is desirable. We describe here a new method using a device comprised of microfluidic channels and a distributed pneumatic pump (BioFlux) that provides fluid flow to 96 individual biofilms. This device allows fine control of continuous or intermittent fluid flow over a broad range of flow rates, and the use of a standard well plate format provides compatibility with plate readers. We show that use of green fluorescent protein (GFP)-expressing bacteria, staining with propidium iodide, and measurement of fluorescence with a plate reader permit rapid and accurate determination of biofilm viability. The biofilm viability measured with the plate reader agreed with that determined using plate counts, as well as with the results of fluorescence microscope image analysis. Using BioFlux and the plate reader, we were able to rapidly screen the effects of several antimicrobials on the viability of Pseudomonas aeruginosa PAO1 flow biofilms.Bacterial biofilms are surface-attached communities that are encased in a polymeric matrix, which exhibit a high degree of resistance to antimicrobial agents and the host immune system (12, 16). This makes them medically important; diseases with a biofilm component are chronic and difficult to eradicate. Examples of such diseases are cystitis (1), endocarditis (31), cystic fibrosis (35), and middle-ear (17) and indwelling medical device-associated (20) infections. Biofilms also play important environmental roles in, for example, wastewater treatment (38), bioremediation (29, 30), biofouling (7), and biocorrosion (2). Better control of biofilms requires elucidation of the molecular basis of their superior resistance (by identifying resistance-compromised mutants) and identification of compounds with antibiofilm activity. While our understanding of these aspects of biofilms has increased (11, 15, 25-27, 36), further work, including development of accurate high-throughput (HTP) methods for screening biofilm viability, is needed.Two major biofilm models are studied in the laboratory, biofilms grown without a continuous flow of fresh medium and biofilms grown with a continuous flow of fresh medium; examples of these two models are microtiter well biofilms and flow cell biofilms, respectively. Methods have been developed for HTP screening of the viability of static biofilms (6, 28, 32, 33), but there are no methods for HTP screening of flow biofilms. The latter biofilms are typically grown in flow cells, which have to be examined individually to determine viability and thus cannot be used for rapid screening. An HTP screening method for flow biofilms is desirable, as these biofilms more closely approximate natural biofilms and can differ from static biofilms evidently due to hydrodynamic influences on cell signaling (22, 34). For example, the ability of rpoS-deficient Escherichia coli (lacking σS) to form flow biofilms is impaired, but its capacity to form biofilms under static conditions is enhanced (18).We describe here a new application of a recently developed device (8-10, 13), the “BioFlux” device consisting of microfluidic channels for biofilm growth. Other microfluidic devices have recently been used for biofilm formation (14, 19, 21, 23), but none of them has been used for HTP screening. The BioFlux device permits rapid measurement of the fluorescence of flow biofilms with a plate reader, which permits initial HTP screening of the viability of such biofilms.  相似文献   

19.
The biofilm matrix contributes to the chemistry, structure, and function of biofilms. Biofilm-derived membrane vesicles (MVs) and DNA, both matrix components, demonstrated concentration-, pH-, and cation-dependent interactions. Furthermore, MV-DNA association influenced MV surface properties. This bears consequences for the reactivity and availability for interaction of matrix polymers and other constituents.The biofilm matrix contributes to the chemistry, structure, and function of biofilms and is crucial for the development of fundamental biofilm properties (46, 47). Early studies defined polysaccharides as the matrix component, but proteins, lipids, and nucleic acids are all now acknowledged as important contributors (7, 15). Indeed, DNA has emerged as a vital participant, fulfilling structural and functional roles (1, 5, 6, 19, 31, 34, 36, 41, 43, 44). The phosphodiester bond of DNA renders this polyanionic at a physiological pH, undoubtedly contributing to interactions with cations, humic substances, fine-dispersed minerals, and matrix entities (25, 41, 49).In addition to particulates such as flagella and pili, membrane vesicles (MVs) are also found within the matrices of gram-negative and mixed biofilms (3, 16, 40). MVs are multifunctional bilayered structures that bleb from the outer membranes of gram-negative bacteria (reviewed in references 4, 24, 27, 28, and 30) and are chemically heterogeneous, combining the known chemistries of the biofilm matrix. Examination of biofilm samples by transmission electron microscopy (TEM) has suggested that matrix material interacts with MVs (Fig. (Fig.1).1). Since MVs produced in planktonic culture have associated DNA (11, 12, 13, 20, 21, 30, 39, 48), could biofilm-derived MVs incorporate DNA (1, 39, 40, 44)?Open in a separate windowFIG. 1.Possible interactions between matrix polymers and particulate structures. Shown is an electron micrograph of a thin section through a P. aeruginosa PAO1 biofilm. During processing, some dehydration occurred, resulting in collapse of matrix material into fibrillate arrangements (black filled arrows). There is a suggestion of interactions occurring with particulate structures such as MVs (hollow white arrow) and flagella (filled white arrows) (identified by the appearance and cross-dimension of these highly ordered structures when viewed at high magnification), which was consistently observed with other embedded samples and also with whole-mount preparations of gently disrupted biofilms (data not shown). The scale bar represents 200 nm.  相似文献   

20.
The occurrence of high concentrations of extracellular DNA (eDNA) in the extracellular matrices of biofilms plays an important role in biofilm formation and development and possibly in horizontal gene transfer through natural transformation. Studies have been conducted to characterize the nature of eDNA and its potential function in biofilm development, but it is difficult to extract eDNA from the extracellular matrices of biofilms without any contamination from genomic DNA released by cell lysis during the extraction process. In this report, we compared several different extraction methods in order to obtain highly pure eDNA from different biofilm samples. After different extraction methods were explored, it was concluded that using chemical treatment or enzymatic treatment of biofilm samples may obtain larger amounts of eDNA than using the simple filtration method. There was no detectable cell lysis when the enzymatic treatment methods were used, but substantial cell lysis was observed when the chemical treatment methods were used. These data suggest that eDNA may bind to other extracellular polymers in the biofilm matrix and that enzymatic treatment methods are effective and favorable for extracting eDNA from biofilm samples. Moreover, randomly amplified polymorphic DNA analysis of eDNA in Acinetobacter sp. biofilms and Acinetobacter sp. genomic DNA and DNA sequencing analysis revealed that eDNA originated from genomic DNA but was not structurally identical to the genomic DNA.A biofilm is a well-organized community of microorganisms that adheres to surfaces and is embedded in the slimy extracellular polymeric substances (EPSs). EPSs are a complex mixture composed of high-molecular-mass polymers (>10,000 Da) generated by the bacterial cells, cell lysis and hydrolysis products, and organic matter adsorbed from the substrate. EPSs are involved in the establishment of stable arrangements of microorganisms in biofilms (40), and it recently was found that extracellular DNA (eDNA) is one of the major components of EPSs (7, 31). eDNA plays a very important role in biofilm development (39), and it is believed to be involved in providing substrates for sibling cells, maintaining the three-dimensional structure of biofilms, and enhancing the exchange of genetic materials (18, 31). eDNA has also been found to be accumulated in cultures of several bacterial species and has been postulated as being released by bacterial cells (11, 15, 21, 30). Although it is commonly accepted that eDNA is released mainly from cell lysis (11, 23, 24, 28, 34, 41), several studies have revealed that some other active secretion mechanisms may exist (1, 6, 11, 27). Recent evidence, however, indicates the possibility that eDNA is secreted actively via transport vesicles for the purpose of creating the biofilm matrix (39). Bockelmann et al. found that eDNA formed a defined, network-like spatial structure in the biofilm of an aquatic bacterium and identified that eDNA was not completely identical to genomic DNA by using randomly amplified polymorphic DNA (RAPD) and restriction endonuclease analyses (3). By using RAPD analysis, principal-components analysis, and terminal restriction fragment length polymorphism analysis, Steinberger and Holden (33) also characterized eDNA in single- and multiple-species unsaturated biofilm and found that it was different from genomic DNA. However, research is still needed to elucidate the role of eDNA in biofilm structures and in the development and origins of eDNA. In order to further investigate these questions, it is important to extract most of the eDNA of high purity in the biofilm matrix and separate eDNA from other components in the EPSs and from the genomic DNA released during the extraction process. Several methods, such as high-speed centrifugation (2, 33) and membrane filtration (3), have been used to isolate eDNA from biofilm samples. However, these methods may isolate only a portion of the eDNA from biofilm samples.EPSs are composed mainly of high-molecular-weight compounds, including polysaccharides, proteins, and amphiphilic polymers (19, 20), that are secreted by microorganisms into their environment (32). The majority of proteins in the EPSs are bridged by divalent ions, including Ca2+ and Mg2+, and a small fraction of carbohydrates and nucleic acids are linked to these divalent ions. Under neutral conditions, the carboxyl of protein would become ionized and negative. Through ion interaction, the divalent ions bridge the protein and the cells. In addition, eDNA may be physically or chemically associated with extracellular proteins, polysaccharides, and other polymers in the EPS matrix. The structural assemblage of proteins and polysaccharides in the complex matrix of the EPS might hinder the liberating eDNA from the EPS matrix. Therefore, it is difficult to release eDNA and other materials from the EPS matrix by only vortexing or homogenizing. Additionally, it is necessary to degrade certain components of EPSs in the biofilm matrix in order to release eDNAs that may bind to these compounds.In this study, the following extractants were chosen to treat biofilm samples for isolation of eDNA from Acinetobacter sp. strain AC811 biofilm: EDTA and cation-exchange resin (CER) (16), which both have the ability to remove cations from the EPS matrix; sodium dodecyl sulfate (SDS) and NaOH, which are strong denaturants and are used frequently for EPS extraction from various pure and mixed cultures (17, 29); and N-glycanase (glycoprotein degradation hydrolase) (35), dispersin B (biofilm-dispersing glycoside hydrolase) (25), and proteinase K (protein hydrolase). We evaluated the efficiencies of these treatments and their impacts on the quantity and quality of eDNA extracted, and we propose that eDNA may bind to other extracellular polymers in the Acinetobacter biofilm matrix, based on the release of eDNA from the biofilm matrix after such treatments.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号