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.
Scanning confocal laser microscopy (SCLM) and fluorescent molecular probes were used to evaluate the effect of the fluoroquinolone fleroxacin on the architecture of established Pseudomonas fluorescens biofilms. Control P. fluorescens biofilms were heterogeneous, consisting of cell aggregates extending from the attachment surface to maximum measured depths of ~90 μm (mean biofilm depth at 72 h, 42 ± 28 μm) and penetrated by an array of channels. In contrast, fleroxacin-treated biofilms were less deep (mean biofilm depth at 72 h, 29 ± 8 μm), varied little in depth over large areas, and consisted of a homogeneous distribution of cells. Fleroxacin also caused cells to elongate, with cells located near the biofilm-liquid interface lengthening significantly more than cells located at the attachment surface. By using SCLM, acridine orange, and image analysis it was found that ~59% of cells within fleroxacin-treated biofilms emitted red fluorescence whereas >99% of cells from control biofilms emitted green fluorescence. The fleroxacin-treated cells which emitted red fluorescence were observed to be the population of cells which elongated.  相似文献   

2.
Nanoparticle uptake and distribution to solid tumors are limited by reticuloendothelial system systemic filtering and transport limitations induced by irregular intra-tumoral vascularization. Although vascular enhanced permeability and retention can aid targeting, high interstitial fluid pressure and dense extracellular matrix may hinder local penetration. Extravascular diffusivity depends upon nanoparticle size, surface modifications, and tissue vascularization. Gold nanoparticles functionalized with biologically-compatible layers may achieve improved uptake and distribution while enabling cytotoxicity through synergistic combination of chemotherapy and thermal ablation. Evaluation of nanoparticle uptake in vivo remains difficult, as detection methods are limited. We employ hyperspectral imaging of histology sections to analyze uptake and distribution of phosphatidylcholine-coated citrate gold nanoparticles (CGN) and silica-gold nanoshells (SGN) after tail-vein injection in mice bearing orthotopic pancreatic adenocarcinoma. For CGN, the liver and tumor showed 26.5±8.2 and 23.3±4.1 particles/100μm2 within 10μm from the nearest source and few nanoparticles beyond 50μm, respectively. The spleen had 35.5±9.3 particles/100μm2 within 10μm with penetration also limited to 50μm. For SGN, the liver showed 31.1±4.1 particles/100μm2 within 10μm of the nearest source with penetration hindered beyond 30μm. The spleen and tumor showed uptake of 22.1±6.2 and 15.8±6.1 particles/100μm2 within 10μm, respectively, with penetration similarly hindered. CGH average concentration (nanoparticles/μm2) was 1.09±0.14 in the liver, 0.74±0.12 in the spleen, and 0.43±0.07 in the tumor. SGN average concentration (nanoparticles/μm2) was 0.43±0.07 in the liver, 0.30±0.06 in the spleen, and 0.20±0.04 in the tumor. Hyperspectral imaging of histology sections enables analysis of phosphatidylcholine-coated gold-based nanoparticles in pancreatic tumors with the goal to improve nanotherapeutic efficacy.  相似文献   

3.
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 microm (titanium) and 107.4 +/- 28.5 microm (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   

4.
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   

5.
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   

6.
We studied the laser ablation of laboratory-developed biofilm on titanium and glass surfaces. Specifically, Pseudoalteromonas carrageenovora, a marine biofilm forming bacterium was used to generate laboratory biofilm. Two fluences, 0.05 and 0.1 J/cm(2) and three durations of irradiation, 30 s, 5 min, and 10 min were tested using an Nd;YAG laser of 532 nm wavelength (in the green light area). Nonirradiated coupons with biofilm served as control. The biofilm removal efficiency increased with the increase in laser fluence and duration of irradiation. The maximum biofilm area cover on control coupons of glass and titanium was 62.5 and 76.0%, respectively. Upon irradiation with fluence 0.1 J/cm(2) for the very short duration of 30 s, this reduced to 5.6 and 12.4% and at 10 min to 2.17 and 0.7% on glass and titanium coupons, respectively, while the controls did not show any reductions (62.5 and 76.0% respectively, for glass and titanium coupons). The biofilm TRC (Total Resuscitated Cells) reduction during this period was even more prominent than the area cover, indicating that the remaining biofilm portions on coupons after irradiation were largely composed of dead bacterial cells. The TRC in the irradiation chamber medium for short durations of irradiation showed a significant increase, indicating that the laser irradiation removed live bacteria from the biofilm. The re-growth of the resuscitated cells showed they could grow like the control cells but with a significant lag. The laser's efficiency in the removal of biofilm was better seen on titanium coupons than on glass. Our results showed that a low-power pulsed laser irradiation could be used to remove biofilm formed on hard surfaces.  相似文献   

7.
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   

8.
Penetration of Rhizopus oligosporus into Soybeans in Tempeh   总被引:1,自引:0,他引:1       下载免费PDF全文
Histological observations were made on the penetration of hyphae of Rhizopus oligosporus into soybean cotyledons in tempeh, an Indonesian soybean food. Hyphal penetrations averaged one per 1,400 μm2 (±390 μm2) on the curved (outer) cotyledon surface and one per 1,010 μm2 (±340 μm2) on the flat (inner) one. Hyphae infiltrated to a depth of 742 μm, or about 25% of the average width of a soybean cotyledon. This previously unreported degree of penetration offers partial explanation for the rapid physical and chemical changes in soybeans during tempeh fermentation.  相似文献   

9.
Aerosols of microorganisms were tested for particle size by use of an Andersen sampler. Mycoplasma aerosols had an average count median diameter (CMD) of 2.1 ± 0.5 μ. Staphylococcus aureus L forms gave an average CMD of 4.6 ± 1.7 μ; the diphtheroid L form, a CMD of 3.4 ± 0.3 μ. Escherichia coli had a CMD of 5.4 ± 2.5 μ; Neisseria sicca, 3.3 ± 0.5 μ; N. meningitidis, 3.4 ± 0.2 μ. S. aureus ATCC 6538, the parent strain of the L form, yielded a CMD of 3.9 ± 1.2 μ. Candida albicans gave an average CMD of 5.9 ± 1.4 μ. All organisms tested survived aerosolizing and could be recovered in viable form for at least 1 hr. Ultraviolet radiation at 2,537 A destroyed the bacteria and mycoplasmas instantaneously, and destroyed 87% of the L forms of S. aureus, 69% of the diphtheroid L form, and 98% of the C. albicans cells. After irradiation, viable particles of the L form and C. albicans aerosols were consistently larger, indicating that clumping led to survival. Submicron size particles were found in aerosols of all species tested except C. albicans.  相似文献   

10.
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   

11.
The resistance of Escherichia coli O157:H7 strains ATCC 43895-, 43895-EPS (an exopolysaccharide [EPS]-overproducing mutant), and ATCC 43895+ (a curli-producing mutant) to chlorine, a sanitizer commonly used in the food industry, was studied. Planktonic cells of strains 43895-EPS and/or ATCC 43895+ grown under conditions supporting EPS and curli production, respectively, showed the highest resistance to chlorine, indicating that EPS and curli afford protection. Planktonic cells (ca. 9 log10 CFU/ml) of all strains, however, were killed within 10 min by treatment with 50 μg of chlorine/ml. Significantly lower numbers of strain 43895-EPS, compared to those of strain ATCC 43895-, attached to stainless steel coupons, but the growth rate of strain 43895-EPS on coupons was not significantly different from that of strain ATCC 43895-, indicating that EPS production did not affect cell growth during biofilm formation. Curli production did not affect the initial attachment of cells to coupons but did enhance biofilm production. The resistance of E. coli O157:H7 to chlorine increased significantly as cells formed biofilm on coupons; strain ATCC 43895+ was the most resistant. Population sizes of strains ATCC 43895+ and ATCC 43895- in biofilm formed at 12°C were not significantly different, but cells of strain ATCC 43895+ showed significantly higher resistance than did cells of strain ATCC 43895-. These observations support the hypothesis that the production of EPS and curli increase the resistance of E. coli O157:H7 to chlorine.  相似文献   

12.

Background

Obesity is associated with diastolic dysfunction, lower maximal myocardial blood flow, impaired myocardial metabolism and increased risk of heart failure. We examined the association between obesity, left ventricular filling pressure and myocardial structure.

Methods

We performed histological analysis of non-ischemic myocardium from 57 patients (46 men and 11 women) undergoing coronary artery bypass graft surgery who did not have previous cardiac surgery, myocardial infarction, heart failure, atrial fibrillation or loop diuretic therapy.

Results

Non-obese (body mass index, BMI, ≤30 kg/m2, n=33) and obese patients (BMI >30 kg/m2, n=24) did not differ with respect to myocardial total, interstitial or perivascular fibrosis, arteriolar dimensions, or cardiomyocyte width. Obese patients had lower capillary length density (1145±239, mean±SD, vs. 1371±333 mm/mm3, P=0.007) and higher diffusion radius (16.9±1.5 vs. 15.6±2.0 μm, P=0.012), in comparison with non-obese patients. However, the diffusion radius/cardiomyocyte width ratio of obese patients (0.73±0.11 μm/μm) was not significantly different from that of non-obese patients (0.71±0.11 μm/μm), suggesting that differences in cardiomyocyte width explained in part the differences in capillary length density and diffusion radius between non-obese and obese patients. Increased BMI was associated with increased pulmonary capillary wedge pressure (PCWP, P<0.0001), and lower capillary length density was associated with both increased BMI (P=0.043) and increased PCWP (P=0.016).

Conclusions

Obesity and its accompanying increase in left ventricular filling pressure were associated with lower coronary microvascular density, which may contribute to the lower maximal myocardial blood flow, impaired myocardial metabolism, diastolic dysfunction and higher risk of heart failure in obese individuals.  相似文献   

13.
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   

14.
Pseudomonas aeruginosa strain NB1 uses chloromethane (CM) as its sole source of carbon and energy under nitrate-reducing and aerobic conditions. The observed yield of NB1 was 0.20 (±0.06) (mean ± standard deviation) and 0.28 (±0.01) mg of total suspended solids (TSS) mg of CM−1 under anoxic and aerobic conditions, respectively. The stoichiometry of nitrate consumption was 0.75 (±0.10) electron equivalents (eeq) of NO3 per eeq of CM, which is consistent with the yield when it is expressed on an eeq basis. Nitrate was stoichiometrically converted to dinitrogen (0.51 ± 0.05 mol of N2 per mol of NO3). The stoichiometry of oxygen use with CM (0.85 ± 0.21 eeq of O2 per eeq of CM) was also consistent with the aerobic yield. Stoichiometric release of chloride and minimal accumulation of soluble metabolic products (measured as chemical oxygen demand) following CM consumption, under anoxic and aerobic conditions, indicated complete biodegradation of CM. Acetylene did not inhibit CM use under aerobic conditions, implying that a monooxygenase was not involved in initiating aerobic CM metabolism. Under anoxic conditions, the maximum specific CM utilization rate (k) for NB1 was 5.01 (±0.06) μmol of CM mg of TSS−1 day−1, the maximum specific growth rate (μmax) was 0.0506 day−1, and the Monod half-saturation coefficient (Ks) was 0.067 (±0.004) μM. Under aerobic conditions, the values for k, μmax, and Ks were 10.7 (±0.11) μmol of CM mg of TSS−1 day−1, 0.145 day−1, and 0.93 (±0.042) μM, respectively, indicating that NB1 used CM faster under aerobic conditions. Strain NB1 also grew on methanol, ethanol, and acetate under denitrifying and aerobic conditions, but not on methane, formate, or dichloromethane.  相似文献   

15.
Calcium is a ubiquitous second messenger in urinary bladder smooth muscle (UBSM). In this study, small discrete elevations of intracellular Ca2+, referred to as Ca2+ sparklets have been detected in an intact detrusor smooth muscle electrical syncytium using a TIRF microscopy Ca2+ imaging approach. Sparklets were virtually abolished by the removal of extracellular Ca2+ (0.035±0.01 vs. 0.23±0.07 Hz/mm2; P<0.05). Co-loading of smooth muscle strips with the slow Ca2+ chelator EGTA-AM (10 mM) confirmed that Ca2+ sparklets are restricted to the cell membrane. Ca2+ sparklets were inhibited by the calcium channel inhibitors R-(+)-Bay K 8644 (1 μM) (0.034±0.02 vs. 0.21±0.08 Hz/mm2; P<0.05), and diltiazem (10 μM) (0.097±0.04 vs. 0.16±0.06 Hz/mm2; P<0.05). Ca2+ sparklets were unaffected by inhibition of P2X1 receptors α,β-meATP (10 μM) whilst sparklet frequencies were significantly reduced by atropine (1 μM). Ca2+ sparklet frequency was significantly reduced by PKC inhibition with Gö6976 (100 nM) (0.030±0.01 vs. 0.30±0.1 Hz/mm2; P<0.05), demonstrating that Ca2+ sparklets are PKC dependant. In the presence of CPA (10 μM), there was no apparent change in the overall frequency of Ca2+ sparklets, although the sparklet frequencies of each UBSM became statistically independent of each other (Spearman''s rank correlation 0.2, P>0.05), implying that Ca2+ store mediated signals regulate Ca2+ sparklets. Under control conditions, inhibition of store operated Ca2+ entry using ML-9 (100 μM) had no significant effect. Amplitudes of Ca2+ sparklets were unaffected by any agonists or antagonists, suggesting that these signals are quantal events arising from activation of a single channel, or complex of channels. The effects of CPA and ML-9 suggest that Ca2+ sparklets regulate events in the cell membrane, and contribute to cytosolic and sarcoplasmic Ca2+ concentrations.  相似文献   

16.
Fallisia arabica n. sp. was described from peripheral blood smears of the Skink lizard, Scincus hemprichii from Jazan Province in the southwest of Saudi Arabia. Schizogony and gametogony take place within neutrophils in the peripheral blood of the host. Mature schizont is rosette shaped 17.5 ± 4.1 × 17.0 ± 3.9 μm, with a L/W ratio of 1.03(1.02–1.05) μm and produces 24(18–26) merozoites. Young gametocytes are ellipsoidal, 5.5 ± 0.8 × 3.6 ± 0.5 μm, with a L/W of 1.53(1.44–1.61) μm. Mature macrogametocytes are ellipsoidal, 9.7 ± 1.2 × 7.8 ± 1.0 μm, with a L/W of 1.24(1.21–1.34) μm and microgametocytes are ellipsoidal, 7.0 ± 1.1 × 6.8 ± 0.9 μm. with a L/W of 1.03(1.01–1.10) μm. In comparison to the described Fallisia species, this new taxon has rosette schizonts and is larger than F. dominicensis, in Hispaniola, F. bipocrati, F. poecilopi, in Panama, F. thecadactyli in Venezuela, and F. effusa, F. simplex, F. modesta, in Brazil. F. arabica has fewer merozoites than F. effusa, F. poecilopi, F. thecadactyli and F. siamense in Thailand. This new species has more merozoites than F. dominicensis and F. modesta. All of these species belong to diverse saurian families (Agamidae, Gekkonidae, Polychrotidae, Scincidae and Teiidae) parasitize only thrombocytes or lymphocytes and some species parasitize immature erythroid cells and leucocytes.  相似文献   

17.
18.

Purpose

To investigate the characteristics of macular ganglion cell-inner plexiform layer (GCIPL) thickness profiles associated with ocular dominance.

Setting

Private practice, Seoul, Republic of Korea.

Design

Comparative case-control study.

Methods

Both eyes of 199 participants with no ophthalmic abnormalities were included. Participants were imaged by spectral-domain optical coherence tomography, and underwent dominant eye testing using a hole-in-a-card test (sighting dominance) at the same visit. Macular GCIPL, as well as circumpapillary retinal nerve fiber layer (RNFL) thickness were compared for individual patients, according to ocular dominance.

Results

Ocular dominance occurred predominantly in the right eye (right vs. left: 72.36 vs. 27.60%; P < 0.001). In the comparison of macular GCIPL thickness, the average (81.27±5.01 μm vs. 80.66±6.31 μm in dominant vs. non-dominant eyes), inferonasal (81.39±5.47μm vs. 80.33±6.82μm, and inferior sectors (77.95±6.05μm vs. 76.97±8.15μm) were significantly different between dominant and non-dominant eyes (P = 0.040, 0.005, and 0.032, respectively). Significant predictors of average GCIPL thickness were spherical equivalent (β = 1.37, P<0.001), astigmatic power (β = 1.44, P = 0.009), disc area (β = 3.90, P < 0.001), average RNFL thickness (β = 0.22, P<0.001), average cup-to-disc ratio (β = 5.74, P = 0.002), difference between the inferior and superior quadrant RNFL thicknesses (β = 0.08, P = 0.024), and ocular dominance (β = 2.10, P = 0.020). On multivariate regression analysis, ocular dominance was correlated with average GCIPL thickness after adjusting for potential confounders (β = 1.63, P = 0.048).

Conclusions

Dominant eyes accompanied significantly thicker average macular GCIPL. This information suggests that macular GCIPL thickness may provide an indicator of the relative dominance of an eye.  相似文献   

19.
We previously reported the exquisite preservation of the ultrastructures of virulent Mycobacterium tuberculosis cells processed through cryofixation and rapid freeze substitution. Here, we report the “structome” analysis (i.e., the quantitative three-dimensional structural analysis of a whole cell at the electron microscopic level) of virulent M. tuberculosis using serial ultrathin sections prepared after cryofixation and rapid freeze substitution and analyzed by transmission electron microscopy. Five M. tuberculosis cells, which were contained in the serial ultrathin cross sections encompassing from one end to the other, were cut into 24, 36, 69, 55, and 63 serial ultrathin sections, respectively. On average, the cells were 2.71 ± 1.05 μm in length, and the average diameter of the cell was 0.345 ± 0.029 μm. The outer membrane and plasma membrane surface areas were 3.04 ± 1.33 μm2 and 2.67 ± 1.19 μm2, respectively. The cell, outer membrane, periplasm, plasma membrane, and cytoplasm volumes were 0.293 ± 0.113 fl (= μm3), 0.006 ± 0.003 fl, 0.060 ± 0.021 fl, 0.019 ± 0.008 fl, and 0.210 ± 0.091 fl, respectively. The average total ribosome number was 1,672 ± 568, and the ribosome density was 716.5 ± 171.4/0.1 fl. This is the first report of a structome analysis of M. tuberculosis cells prepared as serial ultrathin sections following cryofixation and rapid freeze substitution and examined by transmission electron microscopy. These data are based on the direct measurement and enumeration of exquisitely preserved single-cell structures in transmission electron microscopy images rather than calculations or assumptions from indirect biochemical or molecular biological data. In addition, these data may explain the slow growth of M. tuberculosis and enhance understanding of the structural properties related to the expression of antigenicity, acid-fastness, and the mechanism of drug resistance, particularly in regard to the ratio of target to drug concentrations.  相似文献   

20.
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   

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