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 46 毫秒
1.
The effect of surface electrochemical polarization on the growth of cells of Pseudomonas fluorescens (ATCC 17552) on gold electrodes has been examined. Potentials positive or negative to the potential of zero charge (PZC) of gold were applied, and these resulted in changes in cell morphology, size at cell division, time to division, and biofilm structure. At −0.2 V (Ag/AgCl-3 M NaCl), cells elongated at a rate of up to 0.19 μm min−1, rendering daughter cells that reached up to 3.8 μm immediately after division. The doubling time for the entire population, estimated from the increment in the fraction of surface covered by bacteria, was 82 ± 7 min. Eight-hour-old biofilms at −0.2 V were composed of large cells distributed in expanded mushroom-like microcolonies that protruded several micrometers in the solution. A different behavior was observed under positive polarization. At an applied potential of 0.5 V, the doubling time of the population was 103 ± 8 min, cells elongated at a lower rate (up to 0.08 μm min−1), rendering shorter daughters (2.5 ± 0.5 μm) after division, although the duplication times were virtually the same at all potentials. Biofilms grown under this positive potential were composed of short cells distributed in a large number of compact microcolonies. These were flatter than those grown at −0.2 V or at the PZC and were pyramidal in shape. Polarization effects on cell growth and biofilm structure resembled those previously reported as produced by changes in the nutritional level of the culture medium.  相似文献   

2.
The architecture of a Sphingomonas biofilm was studied during early phases of its formation, using strain L138, a gfp-tagged derivative of Sphingomonas sp. strain LB126, as a model organism and flow cells and confocal laser scanning microscopy as experimental tools. Spatial and temporal distribution of cells and exopolymer secretions (EPS) within the biofilm, development of microcolonies under flow conditions representing varied Reynolds numbers, and changes in diffusion length with reference to EPS production were studied by sequential sacrificing of biofilms grown in multichannel flow cells and by time-lapse confocal imaging. The area of biofilm in terms of microscopic images required to ensure representative sampling varied by an order of magnitude when area of cell coverage (2 × 105 μm2) or microcolony size (1 × 106 μm2) was the biofilm parameter under investigation. Hence, it is necessary to establish the inherent variability of any biofilm metric one is attempting to quantify. Sphingomonas sp. strain L138 biofilm architecture consisted of microcolonies and extensive water channels. Biomass and EPS distribution were maximal at 8 to 9 μm above the substratum, with a high void fraction near the substratum. Time-lapse confocal imaging and digital image analysis showed that growth of the microcolonies was not uniform: adjacently located colonies registered significant growth or no growth at all. Microcolonies in the biofilm had the ability to move across the attachment surface as a unit, irrespective of fluid flow direction, indicating that movement of microcolonies is an inherent property of the biofilm. Width of water channels decreased as EPS production increased, resulting in increased diffusion distances in the biofilm. Changing hydrodynamic conditions (Reynolds numbers of 0.07, 52, and 87) had no discernible influence on the characteristics of microcolonies (size, shape, or orientation with respect to flow) during the first 24 h of biofilm development. Inherent factors appear to have overriding influence, vis-à-vis environmental factors, on early stages of microcolony development under these laminar flow conditions.  相似文献   

3.
Biofilms can be undesirable, as in those covering medical implants, and beneficial, such as when they are used for waste treatment. Because cohesive strength is a primary factor affecting the balance between growth and detachment, its quantification is essential in understanding, predicting, and modeling biofilm development. In this study, we developed a novel atomic force microscopy (AFM) method for reproducibly measuring, in situ, the cohesive energy levels of moist 1-day biofilms. The biofilm was grown from an undefined mixed culture taken from activated sludge. The volume of biofilm displaced and the corresponding frictional energy dissipated were determined as a function of biofilm depth, resulting in the calculation of the cohesive energy. Our results showed that cohesive energy increased with biofilm depth, from 0.10 ± 0.07 nJ/μm3 to 2.05 ± 0.62 nJ/μm3. This observation was reproducible, with four different biofilms showing the same behavior. Cohesive energy also increased from 0.10 ± 0.07 nJ/μm3 to 1.98 ± 0.34 nJ/μm3 when calcium (10 mM) was added to the reactor during biofilm cultivation. These results agree with previous reports on calcium increasing the cohesiveness of biofilms. This AFM-based technique can be performed with available off-the-shelf instrumentation. It could therefore be widely used to examine biofilm cohesion under a variety of conditions.  相似文献   

4.
The role of oxygen availability in determining the local physiological activity of Pseudomonas aeruginosa growing in biofilms was investigated. Biofilms grown in an ambient-air environment expressed approximately 1/15th the alkaline phosphatase specific activity of planktonic bacteria subjected to the same phosphate limitation treatment. Biofilms grown in a gaseous environment of pure oxygen exhibited 1.9 times the amount of alkaline phosphatase specific activity of air-grown biofilms, whereas biofilms grown in an environment in which the air was replaced with pure nitrogen prior to the inducing treatment did not develop alkaline phosphatase activity. Frozen cross sections of biofilms stained for alkaline phosphatase activity with a fluorogenic stain demonstrated that alkaline phosphatase activity was concentrated in distinct bands adjacent to the gaseous interfaces. These bands were approximately 30 μm thick with biofilms grown in air, 2 μm thick with biofilms grown in pure nitrogen, and 46 μm thick with biofilms grown in pure oxygen. Overall biofilm thickness ranged from approximately 117 to approximately 151 μm. Measurements with an oxygen microelectrode indicated that oxygen was depleted locally within the biofilm and that the oxygen-replete zone was of a dimension similar to that of the biologically active zone, as indicated by alkaline phosphatase induction. These experiments revealed marked spatial physiological heterogeneity within P. aeruginosa biofilms in which active protein synthesis was restricted by oxygen availability to the upper 30 μm of the biofilm. Such physiological heterogeneity has implications for microbial ecology and for understanding the reduced susceptibilities of biofilms to antimicrobial agents.  相似文献   

5.
In Vitro Laser Ablation of Natural Marine Biofilms   总被引:1,自引:0,他引:1       下载免费PDF全文
We studied the efficiency of pulsed low-power laser irradiation of 532 nm from an Nd:YAG (neodymium-doped yttrium-aluminum-garnet) laser to remove marine biofilm developed on titanium and glass coupons. Natural biofilms with thicknesses of 79.4 ± 27.8 μm (titanium) and 107.4 ± 28.5 μm (glass) were completely disrupted by 30 s of laser irradiation (fluence, 0.1 J/cm2). Laser irradiation significantly reduced the number of diatoms and bacteria in the biofilm (paired t test; P < 0.05). The removal was better on titanium than on glass coupons.  相似文献   

6.
Simultaneous production of sulfide and methane by anaerobic sewer biofilms has recently been observed, suggesting that sulfate-reducing bacteria (SRB) and methanogenic archaea (MA), microorganisms known to compete for the same substrates, can coexist in this environment. This study investigated the community structures and activities of SRB and MA in anaerobic sewer biofilms (average thickness of 800 μm) using a combination of microelectrode measurements, molecular techniques, and mathematical modeling. It was seen that sulfide was mainly produced in the outer layer of the biofilm, between the depths of 0 and 300 μm, which is in good agreement with the distribution of SRB population as revealed by cryosection-fluorescence in situ hybridization (FISH). SRB had a higher relative abundance of 20% on the surface layer, which decreased gradually to below 3% at a depth of 400 μm. In contrast, MA mainly inhabited the inner layer of the biofilm. Their relative abundances increased from 10% to 75% at depths of 200 μm and 700 μm, respectively, from the biofilm surface layer. High-throughput pyrosequencing of 16S rRNA amplicons showed that SRB in the biofilm were mainly affiliated with five genera, Desulfobulbus, Desulfomicrobium, Desulfovibrio, Desulfatiferula, and Desulforegula, while about 90% of the MA population belonged to the genus Methanosaeta. The spatial organizations of SRB and MA revealed by pyrosequencing were consistent with the FISH results. A biofilm model was constructed to simulate the SRB and MA distributions in the anaerobic sewer biofilm. The good fit between model predictions and the experimental data indicate that the coexistence and spatial structure of SRB and MA in the biofilm resulted from the microbial types and their metabolic transformations and interactions with substrates.  相似文献   

7.
The spatial distributions of zinc, a representative transition metal, and active biomass in bacterial biofilms were determined using two-photon laser scanning microscopy (2P-LSM). Application of 2P-LSM permits analysis of thicker biofilms than are amenable to observation with confocal laser scanning microscopy and also provides selective excitation of a smaller focal volume with greater depth localization. Thin Escherichia coli PHL628 biofilms were grown in a minimal mineral salts medium using pyruvate as the carbon and energy source under batch conditions, and thick biofilms were grown in Luria-Bertani medium using a continuous-flow drip system. The biofilms were visualized by 2P-LSM and shown to have heterogeneous structures with dispersed dense cell clusters, rough surfaces, and void spaces. Contrary to homogeneous biofilm model predictions that active biomass would be located predominantly in the outer regions of the biofilm and inactive or dead biomass (biomass debris) in the inner regions, significant active biomass fractions were observed at all depths in biofilms (up to 350 μm) using live/dead fluorescent stains. The active fractions were dependent on biofilm thickness and are attributed to the heterogeneous characteristics of biofilm structures. A zinc-binding fluorochrome (8-hydroxy-5-dimethylsulfoamidoquinoline) was synthesized and used to visualize the spatial location of added Zn within biofilms. Zn was distributed evenly in a thin (12 μm) biofilm but was located only at the surface of thick biofilms, penetrating less than 20 μm after 1 h of exposure. The relatively slow movement of Zn into deeper biofilm layers provides direct evidence in support of the concept that thick biofilms may confer resistance to toxic metal species by binding metals at the biofilm-bulk liquid interface, thereby retarding metal diffusion into the biofilm (G. M. Teitzel and M. R. Park, Appl. Environ. Microbiol. 69:2313-2320, 2003).  相似文献   

8.
A combination of fluorescence in situ hybridization, microprofiles, denaturing gradient gel electrophoresis of PCR-amplified 16S ribosomal DNA fragments, and 16S rRNA gene cloning analysis was applied to investigate successional development of sulfate-reducing bacteria (SRB) community structure and in situ sulfide production activity within a biofilm growing under microaerophilic conditions (dissolved oxygen concentration in the bulk liquid was in the range of 0 to 100 μM) and in the presence of nitrate. Microelectrode measurements showed that oxygen penetrated 200 μm from the surface during all stages of biofilm development. The first sulfide production of 0.32 μmol of H2S m−2 s−1 was detected below ca. 500 μm in the 3rd week and then gradually increased to 0.70 μmol H2S m−2 s−1 in the 8th week. The most active sulfide production zone moved upward to the oxic-anoxic interface and intensified with time. This result coincided with an increase in SRB populations in the surface layer of the biofilm. The numbers of the probe SRB385- and 660-hybridized SRB populations significantly increased to 7.9 × 109 cells cm−3 and 3.6 × 109 cells cm−3, respectively, in the surface 400 μm during an 8-week cultivation, while those populations were relatively unchanged in the deeper part of the biofilm, probably due to substrate transport limitation. Based on 16S rRNA gene cloning analysis data, clone sequences that related to Desulfomicrobium hypogeium (99% sequence similarity) and Desulfobulbus elongatus (95% sequence similarity) were most frequently found. Different molecular analyses confirmed that Desulfobulbus, Desulfovibrio, and Desulfomicrobium were found to be the numerically important members of SRB in this wastewater biofilm.  相似文献   

9.
Candida albicans forms two types of biofilm in RPMI 1640 medium, depending upon the configuration of the mating type locus. In the prevalent a/α configuration, cells form a biofilm that is impermeable, impenetrable by leukocytes, and fluconazole resistant. It is regulated by the Ras1/cyclic AMP (cAMP) pathway. In the a/a or α/α configuration, white cells form a biofilm that is architecturally similar to an a/α biofilm but, in contrast, is permeable, penetrable, and fluconazole susceptible. It is regulated by the mitogen-activated protein (MAP) kinase pathway. The MTL-homozygous biofilm has been shown to facilitate chemotropism, a step in the mating process. This has led to the hypothesis that specialized MTL-homozygous biofilms facilitate mating. If true, then MTL-homozygous biofilms should have an advantage over MTL-heterozygous biofilms in supporting mating. We have tested this prediction using a complementation strategy and show that minority opaque a/a and α/α cells seeded in MTL-homozygous biofilms mate at frequencies 1 to 2 orders of magnitude higher than in MTL-heterozygous biofilms. No difference in mating frequencies was observed between seeded patches of MTL-heterozygous and MTL-homozygous cells grown on agar at 28°C in air or 20% CO2 and at 37°C. Mating frequencies are negligible in seeded patches of both a/α and a/a cells, in contrast to seeded biofilms. Together, these results support the hypothesis that MTL-homozygous (a/a or α/α) white cells form a specialized “sexual biofilm.”  相似文献   

10.
In this study, we examined the long-term development of the overall structural morphology and community composition of a biofilm formed in a model drinking water distribution system with biofilms from 1 day to 3 years old. Visualization and subsequent quantification showed how the biofilm developed from an initial attachment of single cells through the formation of independent microcolonies reaching 30 μm in thickness to a final looser structure with an average thickness of 14.1 μm and covering 76% of the surface. An analysis of the community composition by use of terminal restriction fragment length polymorphisms showed a correlation between the population profile and the age of the sample, separating the samples into young (1 to 94 days) and old (571 to 1,093 days) biofilms, whereas a limited spatial variation in the biofilm was observed. A more detailed analysis with cloning and sequencing of 16S rRNA fragments illustrated how a wide variety of cells recruited from the bulk water initially attached and resulted in a species richness comparable to that in the water phase. This step was followed by the growth of a bacterium which was related to Nitrospira, which constituted 78% of the community by day 256, and which resulted in a reduction in the overall richness. After 500 days, the biofilm entered a stable population state, which was characterized by a greater richness of bacteria, including Nitrospira, Planctomyces, Acidobacterium, and Pseudomonas. The combination of different techniques illustrated the successional formation of a biofilm during a 3-year period in this model drinking water distribution system.  相似文献   

11.
Farnesol is a quorum-sensing molecule that inhibits filamentation in Candida albicans. Both filamentation and quorum sensing are deemed to be important factors in C. albicans biofilm development. Here we examined the effect of farnesol on C. albicans biofilm formation. C. albicans adherent cell populations (after 0, 1, 2, and 4 h of adherence) and preformed biofilms (24 h) were treated with various concentrations of farnesol (0, 3, 30, and 300 μM) and incubated at 37°C for 24 h. The extent and characteristics of biofilm formation were then assessed microscopically and with a semiquantitative colorimetric technique based on the use of 2,3-bis(2-methoxy-4-nitro-5-sulfo-phenyl)-2H-tetrazolium-5-carboxanilide. The results indicated that the effect of farnesol was dependent on the concentration of this compound and the initial adherence time, and preincubation with 300 μM farnesol completely inhibited biofilm formation. Supernatant media recovered from mature biofilms inhibited the ability of planktonic C. albicans to form filaments, indicating that a morphogenetic autoregulatory compound is produced in situ in biofilms. Northern blot analysis of RNA extracted from cells in biofilms indicated that the levels of expression of HWP1, encoding a hypha-specific wall protein, were decreased in farnesol-treated biofilms compared to the levels in controls. Our results indicate that farnesol acts as a naturally occurring quorum-sensing molecule which inhibits biofilm formation, and we discuss its potential for further development and use as a novel therapeutic agent.  相似文献   

12.
A microscopic method for noninvasively visualizing the action of an antimicrobial agent inside a biofilm was developed and applied to describe spatial and temporal patterns of mouthrinse activity on model oral biofilms. Three species biofilms of Streptococcus oralis, Streptococcus gordonii, and Actinomyces naeslundii were grown in glass capillary flow cells. Bacterial cells were stained with the fluorogenic esterase substrate Calcien AM (CAM). Loss of green fluorescence upon exposure to an antimicrobial formulation was subsequently imaged by time-lapse confocal laser scanning microscopy. When an antimicrobial mouthrinse containing chlorhexidine digluconate was administered, a gradual loss of green fluorescence was observed that began at the periphery of cell clusters where they adjoined the flowing bulk fluid and progressed inward over a time period of several minutes. Image analysis was performed to quantify a penetration velocity of 4 μm/min. An enzyme-based antimicrobial formulation led to a gradual, continually slowing loss of fluorescence in a pattern that was qualitatively different from the behavior observed with chlorhexidine. Ethanol at 11.6% had little effect on the biofilm. None of these treatments resulted in the removal of biomass from the biofilm. Most methods to measure or visualize antimicrobial action in biofilms are destructive. Spatial information is important because biofilms are known for their structural and physiological heterogeneity. The CAM staining technique has the potential to provide information about the rate of antimicrobial penetration, the presence of tolerant subpopulations, and the extent of biomass removal effected by a treatment.  相似文献   

13.
Biofilm formed by Staphylococcus aureus significantly enhances antibiotic resistance by inhibiting the penetration of antibiotics, resulting in an increasingly serious situation. This study aimed to assess whether baicalein can prevent Staphylococcus aureus biofilm formation and whether it may have synergistic bactericidal effects with antibiotics in vitro. To do this, we used a clinically isolated strain of Staphylococcus aureus 17546 (t037) for biofilm formation. Virulence factors were detected following treatment with baicalein, and the molecular mechanism of its antibiofilm activity was studied. Plate counting, crystal violet staining, and fluorescence microscopy revealed that 32 μg/mL and 64 μg/mL baicalein clearly inhibited 3- and 7-day biofilm formation in vitro. Moreover, colony forming unit count, confocal laser scanning microscopy, and scanning electron microscopy showed that vancomycin (VCM) and baicalein generally enhanced destruction of biofilms, while VCM alone did not. Western blotting and real-time quantitative polymerase chain reaction analyses (RTQ-PCR) confirmed that baicalein treatment reduced staphylococcal enterotoxin A (SEA) and α-hemolysin (hla) levels. Most strikingly, real-time qualitative polymerase chain reaction data demonstrated that 32 μg/mL and 64 μg/mL baicalein downregulated the quorum-sensing system regulators agrA, RNAIII, and sarA, and gene expression of ica, but 16 μg/mL baicalein had no effect. In summary, baicalein inhibited Staphylococcus aureus biofilm formation, destroyed biofilms, increased the permeability of vancomycin, reduced the production of staphylococcal enterotoxin A and α-hemolysin, and inhibited the quorum sensing system. These results support baicalein as a novel drug candidate and an effective treatment strategy for Staphylococcus aureus biofilm-associated infections.  相似文献   

14.
Cryptosporidium parvum oocysts accumulate on biofilm surfaces. The percentage of oocysts attached to biofilms remained nearly constant while oocysts were supplied to the system but decreased to a new steady-state level once oocysts were removed from the feed. More oocysts attached to summer biofilm cultures than winter biofilm cultures.Cryptosporidium causes a potentially life-threatening gastrointestinal disease. Because conventional water treatment may not effectively target Cryptosporidium, source water monitoring and protection are important to avoid infection outbreaks.Biofilms can accumulate pathogens at densities that are much higher than water column densities, with the potential for pathogen release long after entrapment (5, 13, 15, 19). Biofilms have been identified as a drinking water contamination source (7), causing infections for which the source cannot be identified (4, 6).Several previous studies examined pathogen transport in biofilms using Cryptosporidium parvum oocysts (2, 6, 15, 16) or beads as pathogen surrogates (3, 5, 11, 12). The former studies did not use natural microbial assemblages (2, 16) or quantify oocyst attachment or sloughing (6, 15). The current study provides novel information about C. parvum oocyst attachment to environmental biofilms, including a mass balance analysis to identify the daily number of oocysts that (i) remained in the flowing water or were sloughed from the biofilm and (ii) were attached to the biofilm. We imaged biofilms using scanning confocal laser microscopy, as used in other studies (9, 17, 20, 21), to identify spatial patterns of oocyst attachment.Biofilms were scraped from rocks found in Monocacy Creek (Bethlehem, PA) into 1 liter of creek water in January 2007 (winter biofilm culture) and July 2008 (summer biofilm culture). The biofilm suspension was vacuum filtered through a 6-μm cellulose filter. The filtrate was centrifuged (1,754 × g for 15 min), and the resulting biofilm pellet was resuspended in 1 ml of raw creek water. The cell concentration was quantified by DAPI (4′,6-diamidino-2-phenylindole) staining (14). Cells were split into aliquots (5 × 106 cells each) and stored at −80°C in cryovials containing 30% glycerol.Single-channel flow chambers (length by width by height, 24 mm by 8 mm by 4 mm) with glass coverslips (Stovall Life Science, Inc., Greensboro, NC) were inoculated with 5 × 106 biofilm cells for 24 h before the flow was started. Filter-sterilized creek water was used as the flow medium. A 12-channel peristaltic pump (Ismatec, Glattbrugg, Switzerland) maintained a constant flow of 0.2 mm/s (1).For biofilm imaging, the following two setups were used: (i) 1 × 104 C. parvum oocysts (Iowa isolate; Waterborne, Inc., New Orleans, LA) (all oocysts were used within 3 weeks of shedding) in the influent each day for 3 days and (ii) 3 × 104 C. parvum oocysts added to the influent for the last 24 h of a 3-day flow experiment. Biofilms were imaged with a Zeiss LSM 510 META laser scanning microscope, using an argon laser (458-nm, 477-nm, 488-nm, and 514-nm excitation wavelengths) and a HeNe1 laser (543-nm excitation wavelength). Biofilms were fixed with methanol, blocked using a 1:10 dilution of fetal bovine serum, and stained with 20 μM SYTO 9 (Invitrogen, Molecular Probes, Eugene, OR) (16). C. parvum oocysts in the biofilm were stained with a Cy3-conjugated monoclonal antibody solution specific for Cryptosporidium (Waterborne, Inc.) (16).For the mass balance analysis, C. parvum oocysts (1 × 104 per day for 3 days) were added to 500 ml constantly stirred influent water to keep oocysts in suspension. Influent water was replaced each day. Experiments to quantify sloughing included 2 or 5 additional days with oocyst-free feed water, for a total of 5 or 8 days. Biofilms used for the 3- and 5-day experiments were grown with the winter biofilm culture; biofilms used for the 8-day experiments were grown with the summer biofilm culture.After each 24-hour period, the remaining influent and effluent waters were processed by membrane filtration (MF) and immunomagnetic separation (IMS) to recover the oocysts. On the last day of each experiment, biofilms were scraped from the flow chambers, resuspended in sterile creek water, and also processed by MF and IMS. MF was performed according to the method of Oda et al. (10), using the 3-μm filter only. IMS was performed on the filtrate using the Aureon IMS kit (ImmTech, Inc., New Windsor, MD), and oocysts were dissociated from the magnetic beads with 0.05 M HCl. IMS products were counted by hemocytometry and corrected for MF and IMS processing losses. An average IMS recovery of 65% ± 4.2% standard error (SE) (determined by four trials using 1 × 104 oocysts in deionized water) was used. MF recoveries were consistent within each day but varied between days. Therefore, an MF recovery control was performed each day using 1 × 104 oocysts in 1 liter deionized water to obtain a daily MF correction factor.The mass balance analysis demonstrated that these methods were effective for tracking oocysts throughout the flow system for the experiment''s duration, accounting for all the oocysts within 8% (Table (Table1).1). In a control flow chamber with no biofilm growth (i.e., a clean glass surface), oocyst loss within the system was 1% or less, indicating that very few oocysts attached to any abiotic surface within the flow system. Laboratory biofilms composed of natural microbial assemblages were successfully created, although grazing impacts that would affect biofilm dynamics in the environment were eliminated. The thicknesses of laboratory biofilms (average thickness, 39.6 μm; SD, 4.7 μm; n = 16) were not statistically different (P of 0.17 by independent t test) than those of natural biofilms in Monocacy Creek (average thickness, 35.8 μm; SD, 10.2 μm; n = 36).

TABLE 1.

Mass balance analysis of biofilms grown for 3, 5, and 8 days, with 3-day oocyst dosinga
Biofilm growthNo. of oocysts ± % SE
% of oocysts ± % SE
Avg biofilm thickness ± SE (μm)
InfluentEffluentBiofilmbIn biofilm at end of oocyst dosing (day 3)In biofilm at end of experimentcAccounted for in system
Day 3 (n = 3)1.5 × 104 ± 1.88.6 × 103 ± 9.36.4 × 103 ± 7.143 ± 5.643 ± 5.6100 ± 1.931 ± 6.1
Day 5 (n = 2)1.9 × 104 ± 2.41.8 × 104 ± 3.53.2 × 103 ± 5.340 ± 4.54.8 ± 2.1108 ± 1.037 ± 3.8
Day 8 (n = 2)2.0 × 104 ± 3.11.5 × 104 ± 2.97.6 × 103 ± 1.864 ± 5.028 ± 0.2107 ± 0.942 ± 3.6
Open in a separate windowaData from two or three replicate experiments are presented.bData determined from direct hemacytometer counts of scraped biofilm at the end of the experiment.cCalculated from influent and effluent data [(influent − effluent)/influent].Oocyst attachment location within the biofilm is important for transport dynamics. Oocyst attachment at the biofilm surface may be followed by (i) no transport into the biofilm depth, (ii) burial by biofilm overgrowth, or (iii) transport into the biofilm depth through water channels. In these experiments, oocysts attached to the biofilm surface and were not observed to move to depths or be buried by biofilm overgrowth (Fig. (Fig.1).1). In the 28 biofilms examined, no difference in oocyst attachment location was seen whether oocysts were present in the flow for the entire study duration (n = 14) or whether oocysts were added to the flow on the last study day (n = 14).Open in a separate windowFIG. 1.Top-down projection (A) and cross-sectional view (B) of a summer biofilm culture, with C. parvum oocysts attached at the biofilm surface. Biofilm cells are stained green with SYTO 9; oocysts are stained red with Cy3. The white line in panel A indicates the location of the cross section shown in panel B. Direction of water flow is from right to left. The biofilm is approximately 24-μm thick; oocysts are located 16 μm above the biofilm base.Previous studies (11, 12) also reported that particle attachment and detachment occurred at the biofilm surface. The inner biofilm was denser, with less pore space, while the biofilm surface had more water channels, providing more surface area for particle attachment. The mean pore size in a variety of biofilms was reported as 1.7 to 2.7 μm at the water surface and 0.3 to 0.4 μm at the substrate surface (11), which would restrict larger particle movement, including oocysts (4 to 7 μm).Oocysts became attached to biofilms and rapidly reached a steady state (Fig. (Fig.2),2), as seen in other studies (5, 6). The percentage of oocysts attached to the biofilm remained nearly constant while oocysts were supplied to the system. Once the oocyst supply was removed, the percentage of oocysts in the biofilm decreased to a new steady state. For winter biofilm cultures, the cumulative percentage of oocysts attached to the biofilm at day 3 (i.e., the end of the dosing period; average, 40.0%; SD, 25%; n = 2) was statistically higher (P of 0.003 by independent t test) than the cumulative percentage of oocysts attached to the biofilm at day 5 (average, 4.8%; SD, 1.4%; n = 2). For the summer biofilm cultures, the cumulative percentage of oocysts attached to the biofilm at day 3 (average, 63.7%; SD, 4.5%; n = 2) was also statistically higher (P of 0.01) than the cumulative percentage of oocysts attached to the biofilm at day 5 (average, 33.5%; SD, 1.1%; n = 2). The oocysts that remained in the biofilm at day 5 likely attached to more-stable or sheltered portions of the biofilm that did not slough.Open in a separate windowFIG. 2.Cumulative percentage of oocysts (±SE; n = 2) associated with the biofilm. The cumulative number of oocysts in the biofilm each day was calculated by adding the daily differences between the number of oocysts in the effluent and influent. This number was converted to a percentage by dividing by the cumulative number of influent oocysts. The biofilm accumulation on the last day was determined from the oocysts collected and counted directly from the biofilm, which agreed with the number calculated using the above-described method. Time zero indicates when the flow began; biofilm growth began 24 h earlier by seeding with microbial concentrate at zero flow. The solid black line on the x axis indicates the period of oocyst addition to the inflow. Error bars are smaller than symbols where not visible.The cumulative percentage of oocysts attached to summer biofilm cultures was statistically higher (P of 0.02 and 0.002 at days 3 and 5, respectively, by independent t test) than the cumulative percentage of oocysts attached to the winter biofilm cultures (Fig. (Fig.2).2). In addition, the thickness of summer biofilm cultures (average thickness, 42.1 μm; SD, 4.2 μm; n = 8) was statistically higher (P of 0.03 by independent t test) than that of winter biofilm cultures (average, 37.0 μm; SD, 3.9 μm; n = 8). However, it is unlikely that biofilm thickness explains the increased oocyst attachment to summer biofilm cultures, because all oocysts were observed to attach at the biofilm surface and no oocysts were ever observed within biofilm depths. These observations are in agreement with those of other studies (3, 8, 16) and suggest that other biofilm characteristics (e.g., surface roughness or pore size) may (i) be more important than biofilm thickness for oocyst attachment and (ii) vary with seasonal differences in water chemistry or microbial community caused by water quality differences, such as temperature, pH, or dissolved organic carbon (16).Biofilms are significant reservoirs for oocysts compared to abiotic surfaces (5, 12, 15, 16). Oocysts that remain in the biofilm have important public health implications because they may persist in the biofilm and eventually be released, resulting in potential human exposure.These results confirm that C. parvum oocysts quickly attach to natural microbial biofilms and can be released into the flowing water over time. Oocyst attachment and release dynamics are important for assessing and potentially reducing the risk of human exposure and infection. Although this study used natural stream biofilms, these transport dynamics have important implications for the drinking water industry. Biofilms in the raw water source, represented here by stream biofilms, are linked to drinking water intakes, where any disturbance event can affect water quality. For this reason, a better understanding of the environmental transport of oocysts is important for tracking oocyst contamination, which ultimately affects the drinking water industry. Further investigation is necessary to understand the differences between the summer and winter biofilm cultures as well as the pathogen reservoir that forms in the biofilm.  相似文献   

15.
Restricted Diffusion in Biophysical Systems: Experiment   总被引:3,自引:0,他引:3       下载免费PDF全文
The pulsed-gradient spin echo nuclear magnetic resonance (PGSENMR) technique was used to measure restricted diffusion of water in three types of animal tissue: human blood plasma and red cells; rat and rabbit heart; rat and rabbit liver. Characteristic lengths (L) for restriction of diffusion are estimated from dependence on the measuring time. Limitations on the range of observable restrictive lengths (1.5-15 μm) are discussed.

The decrease in diffusivity due to 1 μm alumina powder (volume fraction = 0.18) in glycerin/water mixtures agrees with the Wang theory assuming spherical particles and no hydration. The characteristic length (L 4 μm) is larger than the particle size (1 μm) or separation (1.8 μm). Comparison of the diffusivities in tissues at short diffusion times with the Wang theory indicates some bound or trapped water.

For packed red blood cells, a restriction (L 2.3 μm) was attributed tothe red cell membrane. A permeability p 0.014 cm/s may be estimated from the decrease in diffusivity. Average values of diffusivity ratio in heart were: 0.36 ± 0.02 for rat; and 0.26 ± 0.03 for rabbit; and in liver: 0.25 ± 0.01 for rat; 0.25 ± .04 for 10-day old rabbit; and 0.195 ± 0.03 for 2-yr old rabbit. A restriction (L 2.7 μm) in rat liver probably results from the mitochondria.

  相似文献   

16.
Enterobacter sakazakii has been reported to form biofilms, but environmental conditions affecting attachment to and biofilm formation on abiotic surfaces have not been described. We did a study to determine the effects of temperature and nutrient availability on attachment and biofilm formation by E. sakazakii on stainless steel and enteral feeding tubes. Five strains grown to stationary phase in tryptic soy broth (TSB), infant formula broth (IFB), or lettuce juice broth (LJB) at 12 and 25°C were examined for the extent to which they attach to these materials. Higher populations attached at 25°C than at 12°C. Stainless steel coupons and enteral feeding tubes were immersed for 24 h at 4°C in phosphate-buffered saline suspensions (7 log CFU/ml) to facilitate the attachment of 5.33 to 5.51 and 5.03 to 5.12 log CFU/cm2, respectively, before they were immersed in TSB, IFB, or LJB, followed by incubation at 12 or 25°C for up to 10 days. Biofilms were not produced at 12°C. The number of cells of test strains increased by 1.42 to 1.67 log CFU/cm2 and 1.16 to 1.31 log CFU/cm2 in biofilms formed on stainless steel and feeding tubes, respectively, immersed in IFB at 25°C; biofilms were not formed on TSB and LJB at 25°C, indicating that nutrient availability plays a major role in processes leading to biofilm formation on the surfaces of these inert materials. These observations emphasize the importance of temperature control in reconstituted infant formula preparation and storage areas in preventing attachment and biofilm formation by E. sakazakii.  相似文献   

17.
When Geobacter sulfurreducens utilizes an electrode as its electron acceptor, cells embed themselves in a conductive biofilm tens of microns thick. While environmental conditions such as pH or redox potential have been shown to change close to the electrode, less is known about the response of G. sulfurreducens to growth in this biofilm environment. To investigate whether respiratory protein abundance varies with distance from the electrode, antibodies against an outer membrane multiheme cytochrome (OmcB) and cytoplasmic acetate kinase (AckA) were used to determine protein localization in slices spanning ∼25 µm-thick G. sulfurreducens biofilms growing on polished electrodes poised at +0.24 V (vs. Standard Hydrogen Electrode). Slices were immunogold labeled post-fixing, imaged via transmission electron microscopy, and digitally reassembled to create continuous images allowing subcellular location and abundance per cell to be quantified across an entire biofilm. OmcB was predominantly localized on cell membranes, and 3.6-fold more OmcB was detected on cells 10–20 µm distant from the electrode surface compared to inner layers (0–10 µm). In contrast, acetate kinase remained constant throughout the biofilm, and was always associated with the cell interior. This method for detecting proteins in intact conductive biofilms supports a model where the utilization of redox proteins changes with depth.  相似文献   

18.
Two-photon laser scanning microscopy (2PLSM) allows fluorescence imaging in thick biological samples where absorption and scattering typically degrade resolution and signal collection of one-photon imaging approaches. The spatial resolution of conventional 2PLSM is limited by diffraction, and the near-infrared wavelengths used for excitation in 2PLSM preclude the accurate imaging of many small subcellular compartments of neurons. Stimulated emission depletion (STED) microscopy is a superresolution imaging modality that overcomes the resolution limit imposed by diffraction and allows fluorescence imaging of nanoscale features. Here, we describe the design and operation of a superresolution two-photon microscope using pulsed excitation and STED lasers. We examine the depth dependence of STED imaging in acute tissue slices and find enhancement of 2P resolution ranging from approximately fivefold at 20 μm to approximately twofold at 90-μm deep. The depth dependence of resolution is found to be consistent with the depth dependence of depletion efficiency, suggesting resolution is limited by STED laser propagation through turbid tissue. Finally, we achieve live imaging of dendritic spines with 60-nm resolution and demonstrate that our technique allows accurate quantification of neuronal morphology up to 30-μm deep in living brain tissue.  相似文献   

19.
Plant defensins are small, cysteine-rich peptides with antifungal activity against a broad range of yeast and fungi. In this study we investigated the antibiofilm activity of a plant defensin from coral bells (Heuchera sanguinea), i.e. HsAFP1. To this end, HsAFP1 was heterologously produced using Pichia pastoris as a host. The recombinant peptide rHsAFP1 showed a similar antifungal activity against the plant pathogen Fusarium culmorum as native HsAFP1 purified from seeds. NMR analysis revealed that rHsAFP1 consists of an α-helix and a triple-stranded antiparallel β-sheet stabilised by four intramolecular disulfide bonds. We found that rHsAFP1 can inhibit growth of the human pathogen Candida albicans as well as prevent C. albicans biofilm formation with a BIC50 (i.e. the minimum rHsAFP1 concentration required to inhibit biofilm formation by 50% as compared to control treatment) of 11.00 ± 1.70 μM. As such, this is the first report of a plant defensin exhibiting inhibitory activity against fungal biofilms. We further analysed the potential of rHsAFP1 to increase the activity of the conventional antimycotics caspofungin and amphotericin B towards C. albicans. Synergistic effects were observed between rHsAFP1 and these compounds against both planktonic C. albicans cells and biofilms. Most notably, concentrations of rHsAFP1 as low as 0.53 μM resulted in a synergistic activity with caspofungin against pre-grown C. albicans biofilms. rHsAFP1 was found non-toxic towards human HepG2 cells up to 40 μM, thereby supporting the lack of a general cytotoxic activity as previously reported for HsAFP1. A structure-function study with 24-mer synthetic peptides spanning the entire HsAFP1 sequence revealed the importance of the γ-core and its adjacent regions for HsAFP1 antibiofilm activity. These findings point towards broad applications of rHsAFP1 and its derivatives in the field of antifungal and antibiofilm drug development.  相似文献   

20.
There is no confocal microscope optimized for single-molecule imaging in live cells and superresolution fluorescence imaging. By combining the swiftness of the line-scanning method and the high sensitivity of wide-field detection, we have developed a, to our knowledge, novel confocal fluorescence microscope with a good optical-sectioning capability (1.0 μm), fast frame rates (<33 fps), and superior fluorescence detection efficiency. Full compatibility of the microscope with conventional cell-imaging techniques allowed us to do single-molecule imaging with a great ease at arbitrary depths of live cells. With the new microscope, we monitored diffusion motion of fluorescently labeled cAMP receptors of Dictyostelium discoideum at both the basal and apical surfaces and obtained superresolution fluorescence images of microtubules of COS-7 cells at depths in the range 085 μm from the surface of a coverglass.  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