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 515 毫秒
1.

A conspicuous bioluminescence during nighttime was reported in an aquaculture farm in the Cochin estuary due to Gonyaulax spinifera bloom on March 20, 2020. In situ measurements on bioluminescence was carried out during nighttime to quantify the response of G. spinifera to various mechanical stimuli. The bioluminescence intensity (BI) was measured using Glowtracka, an advanced single channel sensor, attached to a Conductivity–Temperature–Depth Profiler. In steady environment, without any external stimuli, the bioluminescence generated due to the movement of fishes and shrimps in the water column was not detected by the sensor. However, stimuli such as a hand splash, oar and swimming movements, and a mixer could generate measurable bioluminescence responses. An abundance of?~?2.7?×?106 cells L?1 of G. spinifera with exceptionally high chlorophyll a of 25 mg m?3 was recorded. The BI in response to hand splash was recorded as high as 1.6?×?1011 photons cm?2 s?1. Similarly, BI of?~?1–6?×?1010 photons cm?2 s?1 with a cumulative bioluminescence of?~?2.51?×?1012 photons cm?2 (for 35 s) was recorded when there is a mixer with a constant force of 494 N/800 rpm min?1. The response of G. spinifera was spontaneous with no time lapse between application of stimuli and the bioluminescence response. Interestingly, in natural environment, application of stimulus for longer time periods (10 min) does not lower the bioluminescence intensity due to the replenishment of water thrusted in by the mixer from surrounding areas. We also demonstrated that the bioluminescence intensity decreases with increase in distance from the source of stimuli (mixer) (av. 1.84?×?1010 photons cm?2 s?1 at 0.2 m to av. 0.05?×?1010 photons cm?2 s?1 at 1 m). The BI was highest in the periphery of the turbulent wake generated by the stimuli (av. 3.1?×?1010 photons cm?2 s?1) compared to the center (av. 1.8?×?1010 photons cm?2 s?1). When the stimuli was applied vertically down, the BI decreased from 0.2 m (0.3?×?1010 photons cm?2 s?1) to 0.5 m (0.10?×?1010 photons cm?2 s?1). Our study demonstrates that the BI of G. spinifera increases with increase in mechanical stimuli and decreases with increase in distance from the stimuli.

  相似文献   

2.
The relationship between leaf resistance to water vapour diffusion and each of the factors leaf water potential, light intensity and leaf temperature was determined for leaves on seedling apple trees (Malus sylvestris Mill. cv. Granny Smith) in the laboratory. Leaf cuticular resistance was also determined and transpiration was measured on attached leaves for a range of conditions. Leaf resistance was shown to be independent of water potential until potential fell below — 19 bars after which leaf resistance increased rapidly. Exposure of leaves to CO2-free air extended the range for which resistance was independent of water potential to — 30 bars. The light requirement for minimum leaf resistance was 10 to 20 W m?2 and at light intensities exceeding these, leaf resistance was unaffected by light intensity. Optimum leaf temperature for minimum diffusion resistance was 23 ± 2°C. The rate of change measured in leaf resistance in leaves given a sudden change in leaf temperature increased as the magnitude of the temperature change increased. For a sudden change of 1°C in leaf temperature, diffusion resistance changed at a rate of 0.01 s cm?1 min?1 whilst for a 9°C leaf temperature change, diffusion resistance changed at a rate of 0.1 s cm?1 min?1. Cuticular resistance of these leaves was 125 s cm?1 which is very high compared with resistances for open stomata of 1.5 to 4 s cm?1 and 30 to 35 s cm?1 for stomata closed in the dark. Transpiration was measured in attached apple leaves enclosed in a leaf chamber and exposed to a range of conditions of leaf temperature and ambient water vapour density. Peak transpiration of approximately 5 × 10?6 g cm?2 s?1 occurred at a vapour density gradient from the leaf to the air of 12 to 14 g m?3 after which transpiration declined due presumably to increased stomatal resistance. Leaves in CO2-free air attained a peak transpiration of 11 × 10?6 g cm?2 s?1 due to lower values of leaf resistance in CO2 free air. Transpiration then declined in these leaves due to development of an internal leaf resistance (of up to 2 s cm?1). The internal resistance was masked in leaves at normal CO2 concentrations by the increase in stomatal resistance.  相似文献   

3.
Zoeae of Paralithodes camtschatica were positively phototactic to white light intensities above 1 × 1013 q cm?2 s?1. Negative phototaxis occurred at low (1 × 1012 q cm?2 s?1), but not high intensities (2.2 × 1016q cm?2 s?1). Phototactic response was directly related to light intensity. Zoeae also responded to red, green and blue light. Zoeae were negatively geotactic, but geotaxis was dominated by phototaxis. Horizontal swimming speed of stage 1 zoeae <4 d old was 2.4 ± 0.1 (SE) cms?1 and decreased to 1.7 ± 0.1 cm s?1 in older zoeae (P <0.01). Horizontal swimming speed of stage 2 zoeae was not significantly different from ≥4 d old stage 1 zoeae. Vertical swimming speed, 1.6 ± 0.1 cm s?1, and sinking rate, 0.7 ± 0.1 cm s?1, did not change with ontogeny. King crab zoeae were positively rheotactic and maintained position in horizontal currents less than 1.4 cm s?1. Starvation reduced swimming and sinking rates and phototactic response.  相似文献   

4.
Three rice varieties, cv. Norin 36, cv. Norin 37 and cv. Yubae, were grown in a loam with a 20 cm water-table which gave aerobic conditions to a depth of not less than 15 to 17 cm. Under these conditions Norin 36 grew more vigorously and tillered more frequently than the other two varieties. The rates of oxygen diffusion at 23°C from roots up to 11 cm in length were however appreciably lower for Norin 36 (4.3 × 10?8g · cm?2 of root surface · min?1) than for Norin 37 or Yubae (c. 7.8 × 10?8g). A considerable increase (up to 200 %) in the oxygen diffusion rate (ODR) from the roots occurred if they were cooled to 3°C, and at this temperature differences in ODR between the varieties were not significant. For a purely physical system, because of the decrease in the diffusion coefficient of oxygen in water, and, the increase in oxygen solubility, a drop of c. 20 % in ODR should accompany the above 20°C drop in temperature. A 16 % drop was recorded for artificial ‘roots’ under these conditions. It was concluded that respiratory activity at the higher temperature must have been responsible for the low readings and intervarietal differences observed at 23°C. By increasing the 3°C values by 25 % a mean value of 14.2 × 10?8g · cm?2 of root surface · min?1 was recorded for the three varieties, being the probable ODR at 23°C in the absence of a respiratory factor. Calculations show that respiratory activity removed enough oxygen to reduce the ODR for Norin 36 by more than 9 × 10?8g, and for Norin 37 and Yubae by c. 6.7 × 10?8g · cm?2 of root surface · min?1. Anatomical investigations showed that cortical breakdown was always extensive at 4 to 4.5 cm from the apex of the roots. In some cases however breakdown had not occurred in the basal segment of the root. No opinion could be formed as to whether differences in the amount of cortical breakdown between the varieties might have occasioned the respiratory differences observed. An interesting feature of the root anatomy was the failure of breakdown in those regions of the roots through which lateral roots emerged.  相似文献   

5.
The effect of sulfhydryl oxidase on the rate of disulfide bond formation and polypeptide chain folding in reductively denatured chymotrypsinogen A has been investigated using an immobilized zymogen preparation and a cylindrical quartz flow-through fluorescence cell. Enzymatic oxidation of the 10 sulfhydryl groups in reduced chymotrypsinogen followed first order kinetics at pH 7.0 with an apparent first order rate constant governing sulfhydryl group disappearance of 4.2 × 10?2 min?1. This provides a t12 of 16.3 min for the sulfhydryl oxidase-catalyzed oxidation, whereas 165 min are required for nonenzymatic aerobic oxidation of one-half the sulfhydryl groups. Refolding of the reductively denatured polypeptide chains, monitored by changes in protein fluorescence, did not follow first order kinetics characteristic of a simple two-state mechanism, nor did the return of trypsin activatability. It appears that at least one intermediate must exist in such refolding, in both the uncatalyzed and sulfhydryl oxidase-catalyzed processes. Estimation of the rate constants governing refolding, assuming a single intermediate between the denatured and native states, provided values of 3 × 10?2 min?1 and 7 × 10?3 min?1 for uncatalyzed autoxidation and 4 × 10?2 min?1 and 1.1 × 10?2 min?1 for the sulfhydryl oxidase-catalyzed transition. Thus, enzymic catalysis of disulfide bond formation can lead to apparent catalysis of protein refolding as monitored both by fluorescence and by acquisition of biological function.  相似文献   

6.
Kinetic properties of rat hepatic prolactin receptors   总被引:1,自引:0,他引:1  
Binding of 125I-labelled ovine prolactin to female rat liver membranes underequilibrium conditions showed an apparent Kd of 200 pM, and a Hill coefficient of 1.0. The association rate was second order, with a rate constant K1, of 2.1 × 107, 1.4 × 107, 1.2 × 107 and 4 × 106 M?1. min?1 at 37, 30, 24 and 4° respectively. At 24° there were two components to the dissociation; a faster phase with K?1=1.26 × 10?2. min?1 (T12=55 minutes) and a slower phase with K?1=1.103 × 10?3. min?1. The apparent Kd (from K?1K1) was 1.05 nM for the faster phase and 87.5 pM for the slower phase. These data suggest that there is a conformational change following hormone binding which results in an increased receptor affinity, which effectively prevents release of bound hormone.  相似文献   

7.
J Greve  J Blok 《Biopolymers》1975,14(1):139-154
Measurements of electric birefringence, sedimentation velocity, and biological adsorption rate are used to study the properties of bacteriophage T4B in the presence of excess tryptophan. The adsorption rate determined in borate buffer pH 9 (at 25°C) increases from 0.003 × 10?8 ml min?1 (0.025 M) to 0.130 × 10?8 ml min?1 (0.150 M). The Kerr coefficient, rotational diffusion coefficient, and the sedimentation coefficient of the phage are also dependent on buffer concentration and reach plateau values above 0.12 M given by Ksp = ?(275 ± 18) × 10?9 OD?1 cm2 statvolt?2, D25,w = 133 ± 4 sec?1, and s20,w = 818 ± 11 S. From a comparison of electric birefringence measurements of T4B and T4D it is concluded that T4D and T4B (in the presence of excess tryptophan) exhibit a similar hydrodynamic behavior. The change in physical parameters is solely due to a shift in fiber configuration. At high buffer concentrations the fibers make an angle of approximately 3π/4 with the sheath and the permanent dipole moment is about 200,000 D. This dipole moment is roughly ten times as large as that of a phage particle with nonextended fibers. This difference may be due to a change in hydrodynamic center upon fiber extension or to the presence of positive charges on the fiber tips, or both. At intermediate buffer concentrations the phage population behaves as if it were monodisperse. Probably not all six fibers are extended under such conditions.  相似文献   

8.
K L Wun  W Prins 《Biopolymers》1975,14(1):111-117
Quasi-elastic light scattering as measured by intensity fluctuation (self-beat) spectroscopy in the time domain can be profitably used to follow both the translational diffusion D and the dominant internal flexing mode τint of DNA and its complexes with various histones in aqueous salt solutions. Without histones, DNA is found to have D = 1.6 × 10?8 cm2/sec and τint ? 5 × 10?4 sec in 0.8 M NaCl, 2 M urea at 20°C. Total histone as well as fraction F2A induce supercoiling (D = 2.6 × 10?8 cm2/sec, τint ? 2.8 × 10?4 sec) whereas fraction F1 induces uncoiling (D = 1.0 × 10?8 cm2/sec, τint ? 9.4 × 10?4 sec). Upon increasing the salt concentration to 1.5 M the DNA–histone complex dissociates (D = 1.8 × 10?8 cm2/sec). Upon decreasing the salt concentration to far below 0.8 M, the DNA–histone complex eventually precipitates as a chromatin gel.  相似文献   

9.
Given the increase in the incidence of insulin resistance, obesity, and type 2 diabetes in children and adolescents, it would be of paramount importance to assess quantitative indices of insulin secretion and action during a physiological perturbation, such as a meal or an oral glucose‐tolerance test (OGTT). A minimal model method is proposed to measure quantitative indices of insulin secretion and action in adolescents from an oral test. A 7 h, 21‐sample OGTT was performed in 11 adolescents. The C‐peptide minimal model was identified on C‐peptide and glucose data to quantify indices of β‐cell function: static φs and dynamic φd responsivity to glucose from which total responsivity φ was also measured. The glucose minimal model was identified on glucose and insulin data to estimate insulin sensitivity, SI, which was compared to a reference measure, SIref, provided by a tracer method. Disposition indices, which adjust insulin secretion for insulin action, were then calculated. Indices of β‐cell function were φs = 51.35 ± 8.89 × 10?9min?1, φd = 1,392 ± 258 × 10?9, and φ = 82.09 ± 17.70 × 10?9min?1. Insulin sensitivity was SI = 14.19 ± 2.73 × 10?4, not significantly different from SIref = 14.96 ± 3.04 × 10?4 dl/kg·min per µU/ml, and well correlated: r = 0.98, P < 0.0001, thus indicating that SI can be accurately measured from an oral test. Disposition indices were DIs = 1,040 ± 201 × 10?14 dl/kg/min2 per pmol/l, DId = 33,178 ± 10,720 × 10?14 dl/kg/min per pmol/l, DI = 1,844 ± 522 × 10?14 dl/kg/min2 per pmol/l. Virtually the same minimal model assessment was obtained with a reduced 3 h, 9‐sample protocol. OGTT interpreted with C‐peptide and glucose minimal model has the potential to provide novel insight regarding the regulation of glucose metabolism in adolescents, and to evaluate the effect of obesity and interventions such as diet and exercise.  相似文献   

10.
Ultrasonic telemetry was used to compare post‐release survival and movements of Atlantic sharpnose sharks Rhizoprionodon terraenovae in a coastal area of the north‐east Gulf of Mexico. Ten fish were caught with standardized hook‐and‐line gear during June to October 1999. Atlantic sharpnose sharks were continuously tracked after release for periods of 0·75 to 5·90 h and their positions recorded at a median interval of 9 min. Individual rate of movement was the mean of all distance and time measurements for each fish. Mean ± s.e . individual rate of movement was 0·45 ± 0·06 total lengths per second (LT s?1) and ranged from 0·28 to 0·92 LT s?1 over all fish. Movement patterns did not differ between jaw and internally hooked Atlantic sharpnose sharks. Individual rate of movement was inversely correlated with bottom water temperature at capture (r2 = 0·52, P ≤ 0·05). No consistent direction in movement was detected for Atlantic sharpnose sharks after release, except that they avoided movement towards shallower areas. Capture‐release survival was high (90%), with only one fish not surviving, i.e. this particular fish stopped movement for a period of 10 min. Total rate of movement was total distance over total time (m min?1) for each Atlantic sharpnose shark. Mean total rate of movement was significantly higher immediately after release at 21·5 m min?1 over the first 1·5 h of tracking, then decreased to 11·2 m min?1 over 1·5–6 h, and 7·7 m min?1 over 3–6 h (P ≤ 0·002), which suggested initial post‐release stress but quick recovery from capture. Thus, high survival (90%) and quick recovery indicate that the practice of catch‐and‐release would be a viable method to reduce capture mortality for R. terraenovae.  相似文献   

11.
Hamster liver glutathione peroxidase was purified to homogeneity in three chromatographic steps and with 30% yield. The purified enzyme had a specific activity of approximately 500 μmol cumene hydroperoxide reduced/min/mg of protein at 37 °C, pH 7.6, and 0.25 mm GSH. The enzyme was shown to be a tetramer of indistinguishable subunits, the molecular weight of which was approximately 23,000 as estimated by sodium dodecyl sulfate-polyacrylamide gel electrophoresis. A single isoelectric point of 5.0 was attributed to the active enzyme. Amino acid analysis determined that selenocysteine, identified as its carboxymethyl derivative, was the only form of selenium. One residue of cysteine was found to be present in each glutathione peroxidase subunit. The presence of tryptophan was colorimetrically determined. Pseudo-first-order kinetics of inactivation of the enzyme by iodoacetate was observed at neutral pH with GSH as the only reducing agent. An optimal pH of 8.0 at 37 °C and an activation energy of 3 kcal/mol at pH 7.6 were found. A ter-uni-ping-pong mechanism was shown by the use of an integrated-rate equation. At pH 7.6, the apparent second-order rate constants for reaction of glutathione peroxidase with hydroperoxides were as follows: k1 (t-butyl hydroperoxide), 7.06 × 105 mm min?1; k1 (cumene hydroperoxide), 1.04 × 106 mm?1 min?1; k1 (p-menthane hydroperoxide), 1.2 × 106 mm?1 min?1; k1 (diisopropylbenzene hydroperoxide), 1.7 × 106 mm?1 min?1; k1 (linoleic acid hydroperoxide), 2.36 × 106 mm?1 min?1; k1 (ethyl hydroperoxide), 2.5 × 106 mm?1 min?1; and k1 (hydrogen peroxide), 2.98 × 106 mm?1 min?1. It is concluded that for bulky hydroperoxides, the more hydrophobic the substrate, the faster its reduction by glutathione peroxidase.  相似文献   

12.
Competent Escherichia coli cells are commonly used in bacterial transformation owing to its high permeability for bioorganic macromolecules like plasmid DNA. However, the mass transfer property of competent E. coli cell has not fully investigated. In the present study, mass transfer coefficients of competent and intact E. coli cells in deionized water were evaluated by impedimetric analysis of the release of cytoplasmic compounds. Because competent cells have a higher permeability after chemical treatment, the lumped mass transfer coefficient of a competent cell was approximately 6.5 times larger than that of an intact cell at room temperature. Release of cytoplasmic components was accelerated at an elevated temperature of 42?°C, which is the heat shock temperature used during bacterial transformation. At this elevated temperature, assessed lumped mass transfer coefficients of intact and competent E. coli cells were 9.28?×?10?4?min?1 and 97.10?×?10?4?min?1, respectively. Significant increase in the mass transfer coefficient of the competent cell is caused by cytolysis of cells. The double layer capacitances were also assessed from the electrochemical spectra confirming the enhanced ion release from E. coli cells and rupture of the competent cell under prolonged exposure at the elevated temperature. Impedimetric detection of the ion release with analyses using an equivalent circuit model provides a method to evaluate mass transfer properties of biomolecules.  相似文献   

13.
Larval Atlantic menhaden Brevoortia tyrannus, spawned off North Carolina (U.S.A.) during the winter, undergo cross-shelf transport from the western Gulf Stream edge to coastal bays and estuaries. Variation in water flow direction with depth provides larvae the opportunity to enhance shoreward transport, if they can regulate their depth behaviourally. Temperature, which normally decreases with depth on the continental shelf, is one possible cue for depth regulation. Laboratory-reared larval menhaden of two different ages were exposed to varying relative rates of temperature increases and decreases, which were presented from both above and below the larvae. Temperature decreases from below caused an ascent response in both young and old larvae, but neither responded to this cue from above. The minimum (threshold) relative rates of decrease for initiating ascents were similar (7.9 × 10?2, 10.7 × 10?2° C min?1) for both age larvae as were the minimum absolute amounts of decrease that must occur before a response (0.1, 0.05°C). Young larvae did not respond to a temperature increase, while old larvae ascended regardless of whether the increase was presented from above or below. Threshold relative rates of increase were 8.59 × 10?2°C min?1 from below and 14.79 × 10?2° C min?1 from above. The threshold rates and range of larval speeds during vertical movements were used to calculate vertical temperature gradients that could be perceived. These values were compared to measured gradients in areas inhabited by menhaden larvae. On the continental shelf, detectable temperature gradients appear common for temperature decreases that would occur upon descending and temperature increases upon ascending. However, it is uncommon for larvae to encounter temperature increases upon descending that would initiate an ascent response. These results support the hypothesis that menhaden larvae are capable of using temperature gradients for depth regulation.  相似文献   

14.
Cryptomonas erosa Skuja, a planktonic alga, was grown in batch culture at different combinations of light intensity and temperature, under nutrient saturation. Growth was maximal (1.2 divisions · day?1) at 23.5 C and 0.043 ly · min?1, declining sharply with temperature (0.025 divisions-day?1 at 1 C). With decreasing temperature, the cells showed both light saturation and inhibition at much reduced light intensities. At the same time the compensation light intensity for growth declined towards a minimum of slightly above 0.4 × 10?4 ly · min?1 (~1 ft-c) at 1 C or <0.1 ly · day?1 (PAR). Cell division was more adversely affected by low temperature than carbon uptake, and the resulting excess production of photosynthate was both stored and excreted. Extreme storage of carbohydrates resulted in cell volumes and carbon content ca. 22 and 30 × greater, respectively, than the maxima observed for cells incubated in the dark, whereas, at growth inhibitory light levels, as much as 57% of the total assimilated carbon was excreted. A marked increase in cell pigment was observed at the lowest light levels (<10?3 ly · min?1), at high temperature. The growth response of C. erosa in culture provides insight into the abundance and distribution of cryptomonads and other small algal flagellates in nature.  相似文献   

15.
We have used translational diffusion coefficient measurements and subunit hydrodynamic theory to determine the dimensions and shape of bacterioophage T4D baseplates and tails. The diffusion coefficient of the baseplate, measured by quasielastic laser light scattering (QLS), was determined previously by Wagenknecht and Bloomfield to be D = 8.56 × 10?8 cm2/s. For the tail, we found D = 5.88 × 10?8 cm2/s by QLS, and D = 6.02 × 10?8 cm2/s by combining sedimentation coefficient and molecular weight in the Svedberg equation. These values, which have an uncertainty of ±2.7%, when combined with subunit hydrodynamic theory, enabled us to refine estimates of dimensions obtained by electron microscopy. For the hexagonal baseplate, the vertex-to-vertex distance is about 480 Å, the thickness is 160 Å, and there are six extended short fibers 320-Å long and 40 Å in diameter. When a baseplate of these dimensions is attached to a tail tube-sheath-connector complex 1050-Å long and 240 Å in diameter, the calculated D is 5.93 × 10?8 cm2/s, within 1% of experiment. This combined use of electron microscopy and hydrodynamics, using the former to ascertain shape, and the latter to obtain solution dimensions, is a powerful approach to the structure of biomolecular complexes.  相似文献   

16.
K562 erythroleukaemic cells produced ascorbate when incubated with dehydroascorbic acid. The reduction depended on the number of cells and on the concentration of dehydroascorbic acid. The observed rate consists of a high affinity (apparent) Km 7 μM , Vmax 3·25 pmol min?1 (106 cells)?1 and a low affinity component, which was non-saturable up to 1 mM of DHA (rate increase of 0·1 pmol min?1 (106 cells)?1 (1 μM of DHA?1). The rate was dependent on temperature and was stimulated by glucose and inhibited by phloretin, N-ethylmaleimide, parachloro-mercuribenzoate and thenoyltrifluoroacetone. Although uptake of DHA proceeded at a higher rate than its extracellular reduction, the generation of extracellular ascorbate from DHA cannot be accounted for by intracellular reduction and the release of ascorbate, since the latter was not linear with time and had an initial rate of approximately 3 pmol min?1 (106 cells?1). At a concentration of DHA of 100 μM this is 25 per cent of the observed reduction.  相似文献   

17.
Fluorescence photobleaching recovery techniques have allowed us to measure the lateral mobility of T-independent antigens bound to antigen-specific mouse B cells. The in vitro immunogenicity or tolerogenicity of antigens we have examined, DNP-polymerized flagellin (DNP-POL), and DNP-linear dextran (DNP-DEX), depend upon the antigen dose and epitope density. These factors also determine the mobility of antigen bound to B cell surfaces. For DNP-POL bound to DNP-specific cells, the observed diffusion constants D decrease monotonically with increasing antigen dose and epitope density. Values of D range from 10.4 × 10?11 cm2 sec?1 for DNP0.4-POL at 0.15 μg/ml to 0.8 × 10?11cm2 sec?1for DNP3.5-POL at 30 μg/ml. For receptor-bound DNP-DEX, D depends strongly on antigen epitope density but not observably on antigen concentration. For epitope densities of 1.2 or less, D is close to the value of 21 × 10?11cm2sec?1 observed for single slg receptors. By an epitope density of 4.8, D has fallen to 2.1 × 10?11cm2sec?1. Peak immunogenicities for DNP-POL and DNP-DEX arc observed when antigen- receptor aggregates have mobilities 14-fold and 3-fold lower, respectively, than a single slg molecule.  相似文献   

18.
Yeast biofilms contribute to quality impairment of industrial processes and also play an important role in clinical infections. Little is known about biofilm formation and their treatment. The aim of this study was to establish a multi-layer yeast biofilm model using a modified 3.7 l bench-top bioreactor operated in continuous mode (D = 0.12 h?1). The repeatability of biofilm formation was tested by comparing five bioprocesses with Rhodotorula mucilaginosa, a strain isolated from washing machines. The amount of biofilm formed after 6 days post inoculation was 83 μg cm?2 protein, 197 μg cm?2 polysaccharide and 6.9 × 106 CFU cm?2 on smooth polypropylene surfaces. Roughening the surface doubled the amount of biofilm but also increased its spatial variability. Plasma modification of polypropylene significantly reduced the hydrophobicity but did not enhance cell attachment. The biofilm formed on polypropylene coupons could be used for sanitation studies.  相似文献   

19.
Alkaline hydrolysis and subcritical water degradation were investigated as ex-situ remediation processes to treat explosive-contaminated soils from military training sites in South Korea. The addition of NaOH solution to the contaminated soils resulted in rapid degradation of the explosives. The degradation of explosives via alkaline hydrolysis was greatly enhanced at pH ≥12. Estimated pseudo-first-order rate constants for the alkaline hydrolysis of 2,4-dinitrotoluene (DNT), 2,4,6-trinitrotoluene (TNT) and hexahydro-1,3,5-trinitro-1,3,5-triazine (RDX) in contaminated soil at pH 13 were (9.6?±?0.1)×10?2, (2.2?±?0.1)×10?1, and (1.7?±?0.2)×10?2 min?1, respectively. In the case of subcritical water degradation, the three explosives were completely removed at 200–300°C due to oxidation at high temperatures and pressures. The degradation rate increased as temperature increased. The pseudo-first-order rate constants for DNT, TNT, and RDX at 300°C were (9.4?±?0.8)×10?2, (22.8?±?0.3)×10?2, and (16.4?±?1.0)×10?2, respectively. When the soil-to-water ratio was more than 1:5, the extent of alkaline hydrolysis and subcritical water degradation was significantly inhibited.  相似文献   

20.
Abstract. Metabolic rates of adult Lophopilio palpinalis (Herbst, 1799) (Arachnida, Opiliones, Phalangioidea) and Paranemastoma quadripunctatum (Perty, 1833) (Arachnida, Opiliones, Troguloidea) are measured during rest and activity. Carbon dioxide release during rest is continuous in both species. Mean values at 20 °C are 4.2 µL min−1 g−1 for the males of P. quadripunctatum, 4.1 µL min−1 g−1 for the males of L. palpinalis and 4.7 µL min−1 g−1 for the females of L. palpinalis, thus being significantly higher in the egg-producing females. In L. palpinalis, respiratory quotient at rest is 0.84. Spontaneous walking activity with speeds of 15–30 cm min−1 raises the metabolic rate by up to three-fold in both species. Lophopilio palpinalis is made to undertake constant running on a treadmill with speeds of 60, 72 and 96 cm min−1. Enforced activity causes the animals to raise their metabolic rates by up to five-fold above resting rates. Animals reach a steady state of CO2 release on the treadmill and show a fast t1/2 on-response, indicating aerobic exercise. The minimum cost of locomotion is determined to be 2.5 × 10−3 J cm−1 g−1, thus fitting the predicted values for terrestrial locomotion.  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