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.
The synaptic apparatus in the ventral nucleus of the medial geniculate body (MGBv) of the cat was examined using electron microscopy and stereological methods, which made it possible to measure the synaptic density. Within 7015 µm2 of examined sections, 1586 presynaptic terminal (PST) profiles were found, which corresponds to 226.0·103 PST per 1 mm2 of section surface. The PSP were classified into five groups:RL,RS,F,P, andUT, in accordance with their ultrastructural pattern (dimension of PST profile, dimension and shape of synaptic vesicles, and type of synaptic contact, SC) [18–22]. On the above surface, there were 1012 SC formed by PST of different groups, which corresponds to 144.0·103 SC per 1 mm2 of section surface. TheRL-,RS-,F-,P-, andUT-type PST formed 14.8%, 50.1%, 13.1%, 16.8%, and 5.2% of analyzed SC, respectively. The calculated mean SC numerical density equalled (260.8±54.8)·106 SC per 1 mm3 of fixed MGBv tissue. Among them, 40.2·106 (15.4%) belonged toRL-PST, i.e., to axonal terminals of thecolliculus inferior neurons; 130.2·106 (49.9%) toRS-PST, i.e., mostly to axonal terminals of the auditory cortex neurons; and 33.9·106 (13.0%) toF-PST, i.e., to axons of the GABA-ergic interneurons and neurons of the perigeniculate division of the reticular thalamic nucleus. Group-P PST, i.e., terminal structures of the dendritic arborizations of interneurons, formed 42.7·106 (16.4%) SC per 1 mm3, and 13.8·106 (5.3%) SC belonged toUT-PST, i.e., to terminals of unidentified nature. Among 260.8·106 SC in 1 mm3 of tissue, only 23.8·106 (9% of total number)RL-SC, localized on the relay neurons, are directly involved in the MGBv relay function. All other SC transmit control influences from various structures of the nervous system, and provide adjustment of relay function to the constantly changing environmental conditions and varying status of an orgamism. The mean number of SC, localized on an averaged MGBv relay neuron, was calculated as 9100. Among them, about 1200 SC belong toRL-PST, 5200 SC toRS-PST, 1200 SC toF-PST, 1100 SC toP-PST, and 400 SC toUT-PST.Neirofiziologiya/Neurophysiology, Vol. 27, No. 3, pp. 208–219, May–June, 1995.  相似文献   

2.
Chang YC  Tsai CY  Lin CF  Wang YC  Wang IK  Chung TC 《Anaerobe》2011,17(5):239-245
To investigate the frequency of tetracycline resistance (Tet-R) lactobacilli in pig intestines, a total of 256 pig colons were analyzed and found to contain typical colonies of Tet-R lactic acid bacteria in every sample, ranging from 3.2 × 103 to 6.6 × 105 CFU/cm2. From these samples, a total of 159 isolates of Tet-R lactobacilli were obtained and identified as belonging to 11 species, including Lactobacillus reuteri, Lactobacillus amylovorus, Lactobacillus salivarius, Lactobacillus plantarum, Lactobacillus ruminis, Lactobacillus kefiri, Lactobacillus fermentum, Lactobacillus sakei, Lactobacillus coryniformis, Lactobacillus parabuchneri and Lactobacillus letivazi. Based on the EFSA (2008) breakpoints, all isolates, after MIC analysis, were qualified as Tet-R, from which the significant high Tet-R MIC50 and MIC90 values indicated an ecological distribution of Tet-R lactobacilli mostly with high resistance potency in pig colons. PCR-detection identified 5 tet genes in these isolates, the most predominant one being tet (W), followed by tet (M), (L), (K), and (Q). Their detection rates were 82.0%, 22.5%, 14.4%, 8.1% and 0.9%, respectively. Noteworthily, isolates of the same species carrying identical tet gene(s) usually had a wide different MIC values. Furthermore, strain-subtyping of these isolates by REP-PCR demonstrated a notable genotypic biodiversity % (average = 62%).  相似文献   

3.
Summary To estimate the advantage of the small red blood cells (RBC) of high-altitude camelids for O2 transfer, the kinetics of O2 uptake into and release from the RBC obtained from llama, vicuña and alpaca were investigated at 37°C with a stopped-flow technique. O2 transfer conductance of RBC (G) was estimated from the rate of O2 saturation change and the corresponding O2 pressure difference between medium and hemoglobin. For comparison, O2 kinetics for the RBC of a lowaltitude camelid (dromedary camel) and the pygmy goat were determined and previously measured values for human RBC were used. O2 transfer of RBC was found to be strongly influenced by extracellular diffusion, except with O2 release into dithionite solutions of sufficiently high concentration (>30 mM). TheG values measured in these standard conditions,G st (in mmol · min–1 · Torr–1 · (ml RBC)–1) were: high-altitude camelids, 0.58 (averaged for llama, alpaca and vicuña since there were no significant interspecific differences); camel 0.42; goat, 0.42; man, 0.39. The differences can in part be attributed to expected effects of the size and shape of the RBC (volume, surface area, mean thickness), as well as to the intracellular O2 diffusivity which depends on the concentration of cellular hemoglobin. The highG st of RBC of highaltitude camelids may be considered to enhance O2 transfer in lungs and tissues. But the O2 transfer conductance of blood, , equal toG st multiplied by hematocrit (in mmol · min–1 · Torr–1 · (ml blood)–1), was only slightly higher as compared to other species: 0.20 (llama, alpaca, vicuña), 0.14 (camel), 0.18 (goat), 0.17 (man).Abbreviations DPG 2,3-diphosphoglycerate - G conductance - Hb hemoglobin - RBC red blood cells - percent saturation of hemoglobin  相似文献   

4.
Summary Responses to acute hypoxia were measured in skipjack tuna (Katsuwonus pelamis) and yellowfin tuna (Thunnus albacares) (1–3 kg body weight). Fish were prevented from making swimming movements by a spinal injection of lidocaine and were placed in front of a seawater delivery pipe to provide ram ventilation of the gills. Fish could set their own ventilation volumes by adjusting mouth gape. Heart rate, dorsal and ventral aortic blood pressures, and cardiac output were continuously monitored during normoxia (inhalant water (PO 2>150 mmHg) and three levels of hypoxia (inhalant water PO 2130, 90, and 50 mmHg). Water and blood samples were taken for oxygen measurements in fluids afferent and efferent to the gills. From these data, various measures of the effectiveness of oxygen transfer, and branchial and systemic vascular resistance were calculated. Despite high ventilation volumes (4–71·min-1·kg-1), tunas extract approximately 50% of the oxygen from the inhalant water, in part because high cardiac outputs (115–132 ml·min-1·kg-1) result in ventilation/perfusion conductance ratios (0.75–1.1) close to the theoretically ideal value of 1.0. Therefore, tunas have oxygen transfer factors (ml O2·min-1·mmHg-1·kg-1) that are 10–50 times greater than those of other fishes. The efficiency of oxygen transfer from water in tunas (65%) matches that measured in teleosts with ventilation volumes and order of magnitude lower. The high oxygen transfer factors of tunas are made possible, in part, by a large gill surface area; however, this appears to carry a considerable osmoregulatory cost as the metabolic rate of gills may account for up 70% of the total metabolism in spinally blocked (i.e., non-swimming) fish. During hypoxia, skipjack and yellowfin tunas show a decrease in heart rate and increase in ventilation volume, as do other teleosts. However, in tunas hypoxic bradycardia is not accompanied by equivalent increases, in stroke volume, and cardiac output falls as HR decreases. In both tuna species, oxygen consumption eventually must be maintained by drawing on substantial venous oxygen reserves. This occurs at a higher inhalant water PO2 (between 130 and 90 mmHg) in skipjack tuna than in yellowfin tuna (between 90 and 50 mmHg). The need to draw on venous oxygen reserves would make it difficult to meet the oxygen demand of increasing swimming speed, which is a common response to hypoxia in both species. Because yellowfin tuna can maintain oxygen consumption at a seawater oxygen tension of 90 mmHg without drawing on venous oxygen reserves, they could probably survive for extended periods at this level of hypoxia.Abbreviations BPda, BPva dorsal, ventral aortic blood pressure - C aO2, C vO2 oxygen content of arterial, venous blood - DO2 diffusion capacity - Eb, Ew effectiveness of O2 uptake by blood, and from water, respectively - Hct hematocrit - HR heart rate - PCO2 carbon dioxide tension - P aCO2, P vCO2 carbon dioxide tension of arterial and venous blood, respectively - PO2 oxygen tension - P aO2, P vO2, P iO2, P cO2 oxygen tension of arterial blood, venous blood, and inspired and expired water, respectively - pHa, pHv pH of arterial and venous blood, respectively - Pw—b effective water to blood oxygen partial pressure difference - Pg partial pressure (tension) gradient - cardiac output - R vascular resistance - SV stroke volume - SEM standard error of mean - TO2 transfer factor - U utilization - g ventilation volume - O2 oxygen consumption  相似文献   

5.
Using the impedance cardiography method, heart rate ( c) matched changes on indexed stroke volume (SI) and cardiac output (CI) were compared in subjects engaged in different types of training. The subjects consisted of untrained controls (C), volleyball players (VB) who spent about half of their training time (360 min · week–1) doing anaerobic conditioning exercises and who had a maximal oxygen uptake ( ) 41% higher than the controls, and distance runners (D) who spent all their training time (366 min·week–1) doing aerobic conditioning exercises and who had a 26% higher than VB. The subjects performed progressive submaximal cycle ergometer exercise (10 W·min–1) up to c of 150 beats·min–1. In group C, SI had increased significantly (P<0.05) at c of 90 beats·min–1 ( + 32%) and maintained this difference up to 110 beats·min–1, only to return to resting values on reaching 130 beats·min–1 with no further changes. In group VB, SI peaked (+ 54%) at c of 110 beats·min–1, reaching a value significantly higher than that of group C, but decreased progressively to 22010 of the resting value on reaching 150 beats·min–1. In group D, SI peaked at c of 130 beats·min–1 (+ 54%), reaching a value significantly higher than that of group VB, and showed no significant reduction with respect to this peak value on reaching 150 beats·min–1. As a consequence, the mean CI increase per c unit was progressively higher in VB than in C (+46%) and in D than in VB (+ 105%). It was concluded that thef c value at which SI ceased to increase during incremental exercise was closely related to the endurance component in the training programme.  相似文献   

6.
Desulfotomaculum acetoxidans has been proposed to oxidize acetate to CO2 via an oxidative acetyl-CoA/carbon monoxide dehydrogenase pathway rather than via the citric acid cycle. We report here the presence of the enzyme activities required for the operation of the novel pathway. In cell extracts the following activities were found (values in brackets=specific activities and apparent K m; 1 U·mg-1=1 mol·min-1·mg protein-1 at 37°C): Acetate kinase (6.3 U·mg-1; 2 mM acetate; 2.4 mM ATP); phosphate acetyltransferase (60 U·mg-1, 0.4 mM acetylphosphate; 0.1 mM CoA); carbon monoxide dehydrogenase (29 U·mg-1; 13% carbon monoxide; 1.3 mM methyl viologen); 5,10-methylenetetrahydrofolate reductase (3 U·mg-1, 0.06 mM CH3–FH4); methylenetetrahydrofolate dehydrogenase (3.6 U·mg-1, 0.9 mM NAD, 0.1 mM CH2=FH4); methenyltetrahydrofolate cyclohydrolase (0.3 U·mg-1); formyltetrahydrofolate synthetase (3 U·mg-1, 1.4 mM FH4, 0.4 mM ATP, 13 mM formate); and formate dehydrogenase (10 U·mg-1, 0.4 mM formate, 0.5 mM NAD). The specific activities are sufficient to account for the in vivo acetate oxidation rate of 0.26 U·mg-1.Non-standard abbreviations FH4 Tetrahydrofolate - CHO-FH4 N10-formyltetrahydrofolate - CHFH4 N5,N10-methenyltetrahydrofolate - CH2=FH4 N5,N10-methylenetetrahydrofolate - CH3–FH4 N5-methyltetrahydrofolate - MOPS morpholinopropane sulfonic acid - DTT d,l-1,4-dithiothreitol - TRIS tris-(hydroxymethyl)-aminomethane - Ap5A p1,P5-di(adenosine-5)pentaphosphate - MV methyl viologen  相似文献   

7.
I. Nijs  I. Impens  T. Behaeghe 《Planta》1989,177(3):312-320
The relationship between leaf photosynthetic capacity (p n, max), net canopy CO2- and H2O-exchange rate (NCER and E t, respectively) and canopy dry-matter production was examined in Lollium perenne L. cv. Vigor in ambient (363±30 l· l-1) and elevated (631±43 l·l-1) CO2 concentrations. An open system for continuous and simultaneous regulation of atmospheric CO2 concentration and NCER and E t measurement was designed and used over an entire growth cycle to calculate a carbon and a water balance. While NCERmax of full-grown canopies was 49% higher at elevated CO2 level, stimulation of p n, max was only 46% (in spite of a 50% rise in one-sided stomatal resistance for water-vapour diffusion), clearly indicating the effect of a higher leaf-area index under high CO2 (approx. 10% in one growing period examined). A larger amount of CO2-deficient leaves resulted in higher canopy dark-respiration rates and higher canopy light compensation points. The structural component of the high-CO2 effect was therefore a disadvantage at low irradiance, but a far greater benefit at high irradiance. Higher canopy darkrespiration rates under elevated CO2 level and low irradiance during the growing period are the primary causes for the increase in dry-matter production (19%) being much lower than expected merely based on the NCERmax difference. While total water use was the same under high and low CO2 levels, water-use efficiency increased 25% on the canopy level and 87% on a leaf basis. In the course of canopy development, allocation towards the root system became greater, while stimulation of shoot dry-matter accumulation was inversely affected. Over an entire growing season the root/shoot production ratio was 22% higher under high CO2 concentration.Abbreviations and symbols C350 ambient CO2, 363±30 l·l-1 - C600 high CO2, 631±43 l·l-1 - c a atmospheric CO2 level - c i CO2 concentration in the intracellular spaces of the leaf - Et canopy evapotranspiration - I o canopy light compensation point - NCER canopy CO2-exchange rate - p n leaf photosynthetic rate - PPFD photosynthetic photon flux density - r a leaf boundary-layer resistance - RD canopy dark-respiration rate - r s stomatal resistance - WUE water use efficiency  相似文献   

8.
Summary The function of the caecal bulb, and its adaptation to chronic high- or low-Na+ intake, was investigated by in vivo perfusion of anaesthetised birds. Effects of acute aldosterone injection (125 g·kg–1 body mass) were also measured.Evidence was found for primary active net absorption of Na+, inducing parallel Na-linked absorption of water and Cl and secretion of K+. Around 20–35% of total Cl absorption and K+ secretion were independent of Na+ fluxes, and these components appear to be driven by passive processes with apparent conductances of 6.3×10–3 (G Cl) and 1.1×10–3 (G K) S·cm–2.Acetate (40mM) stimulated Na+ fluxes (8.5–9.9 Eq·cm–2·h–1) and Na-linked water fluxes (27–44 l·cm–2·h–1). Increased coupling ratios (2.9–4.6 l·Eq–1) and other data indicate that these effects may be due to increased osmotic permeabilities of barriers involved in the Na-linked water transfer pathway.Low-Na+ maintenance enhanced EPD (49–69 mV, serosa positive) and all net fluxes:J Na (6.8–11.6);J K (–3.2––4.3);J Cl (4.3–5.6 Eq·cm serosal area–2·h–1);J v (28–43 l·cm–2·h–1) (mucosal-serosal fluxes positive).Acute aldosterone enhancedJ Na (10.8–14.0 Eq·cm–2·h–1) and EPD (54–66 mV) by 3 h after injection, but had no effect on the Na-linked components ofJ K orJ Cl.Abbreviations ECPD, EPD Electrochemical or electrical potential difference - G Cl ,G K ionic conductances (Cl, K+) - J v ,J ion net volume or ion flux rate, mucosa-serosa positive;P d (Cl) diffusive permeability coefficient (of Cl) - SEDM standard error of difference between means  相似文献   

9.
The stability and, consequently, the lifetime of immobilized enzymes (IME) are important factors in practical applications of IME, especially so far as design and operation of the enzyme reactors are concerned. In this paper a model is presented which describes the effect of intraparticle diffusion on time stability behaviour of IME, and which has been verified experimentally by the two-substrate enzymic reaction. As a model reaction the ethanol oxidation catalysed by immobilized yeast alcohol dehydrogenase was chosen. The reaction was performed in the batch-recycle reactor at 303 K and pH-value 8.9, under the conditions of high ethanol concentration and low coenzyme (NAD+) concentration, so that NAD+ was the limiting substrate. The values of the apparent and intrinsic deactivation constant as well as the apparent relative lifetime of the enzyme were calculated.The results show that the diffusional resistance influences the time stability of the IME catalyst and that IME appears to be more stabilized under the larger diffusion resistance.List of Symbols C A, CB, CE mol · m–3 concentration of coenzyme NAD+, ethanol and enzyme, respectively - C p mol · m3 concentration of reaction product NADH - d p mm particle diameter - D eff m2 · s–1 effective volume diffusivity of NAD+ within porous matrix - k d s–1 intrinsic deactivation constant - K A, KA, KB mol · m–3 kinetic constant defined by Eq. (1) - K A x mol · m–3 kinetic constant defined by Eq. (5) - r A mol · m–3 · s–1 intrinsic reaction rate - R m particle radius - R v mol · m–3 · s–1 observed reaction rate per unit volume of immobilized enzyme - t E s enzyme deactivation time - t r s reaction time - V mol · m–3 · s–1 maximum reaction rate in Eq. (1) - V x mol · m–3 · s–1 parameter defined by Eq. (4) - V f m3 total volume of fluid in reactor - w s kg mass of immobilized enzyme bed - factor defined by Eqs. (19) and (20) - kg · m–3 density of immobilized enzyme bed - unstableness factor - effectiveness factor - Thiele modulus - relative half-lifetime of immobilized enzyme Index o values obtained with fresh immobilized enzyme  相似文献   

10.
Specific nitrogenase activity inAzospirillum brasilense ATCC 29145 in surface cultures under air is enhanced from about 50 nmol C2H4·mg protein-1·h-1 to 400 nmol C2H4 by the addition of 1 mM phenol. 0.5 and 2 mM phenol added increase the rate 5-fold and 4-fold. This enhancement effect is observed only between 2 and 3 days after inoculation, with only a small reduction of the growth of the cells by the phenol added. In surface cultures under 1% O2, nitrogenase activity is slightly reduced by the addition of 1–0.01 mM phenol. Utilization of succinate is enhanced during the period of maximum enhancement of nitrogenase activity by 60% by addition of 1 mM phenol. The cells did not produce14CO2 from [U-14C] phenol, neither in surface cultures nor in liquid cultures and less than 0.1% of the phenol was incorporated into the cells. A smaller but significant enhancement of nitrogenase activity by about 100% in surface cultures under air was found withKlebsiella pneumoniae K 11 after addition of 1 mM phenol. However, inRhizobium japonicum 61-A-101 all phenol concentrations above 0.01 mM reduced nitrogenase activity. With 1 mM phenol added activity was reduced to less than 10% with no effect on the growth in the same cultivation system. With thisRhizobium japonicum strain significant quantities of phenol (25 mol in 24 h by 2·1012 cells) were metabolized to14CO2, with phenol as sole carbon source. WithAzospirillum brasilense in liquid culture under 1% and 2% O2 in the gas phase, no enhancement of nitrogenase activity by phenol was noticed.  相似文献   

11.
Summary The influence of the concentration of oxygen on lipase production by the fungus Rhizopus delemar was studied in different fermenters. The effect of oxygen limitation ( 47 mol/l) on lipase production by R. delemar is large as could be demonstrated in pellet and filamentous cultures. A model is proposed to describe the extent of oxygen limitation in pellet cultures. Model estimates indicate that oxygen is the limiting substrate in shake flask cultures and that an optimal inoculum size for oxygen-dependent processes can occur.Low oxygen concentrations greatly negatively affect the metabolism of R. delemar, which could be shown by cultivation in continuous cultures in filamentous growth form (Doptimal=0.086 h-1). Continuous cultivations of R. delemar at constant, low-oxygen concentrations are a useful tool to scale down fermentation processes in cases where a transient or local oxygen limitation occurs.Symbols and Abbreviations CO Oxygen concentration in the gas phase at time = 0 (kg·m-3) - CO 2i Oxygen concentration at the pellet liquid interface (kg·m-3) - CO 2i Oxygen concentration in the bulk (kg·m-3) - D Dilution rate (h-1) - IDO 2 Diffusion coefficient for oxygen (m2·s-1) - dw Dry weight of biomass (kg) - f Conversion factor (rs O 2 to oxygen consumption rate per m3) (-) - k Radial growth rate (m·s-1) - K Constant - kla Volumetric mass transfer coefficient (s-1) - klA Oxygen transfer rate (m-3·s-1) - kl Mass transfer coefficient (m·s-1) - K O 2 Affinity constant for oxygen (mol·m-3) - K w Cotton plug resistance (m-3·s-1) - M Henry coefficient (-) - NV Number of pellets per volume (m-3) - R Radius (m) - RO Radius of oxygen-deficient core (m) - RQ Respiration quotient (mol CO2/mol O2) - rs O 2 Specific oxygen consumption rate per dry weight biomass (kg O2·s-1[kg dw]-1) - rX Biomass production rate (kg·m-3·s-1) - SG Soytone glucose medium (for shake flask experiments) - SG 4 Soytone glucose medium (for tower fermenter and continuous culture experiments) - V Volume of medium (m-3) - X Biomass (dry weight) concentration (kg·m-3) - XR o Biomass concentration within RO for a given X (kg·m-3) - Y O 2 Biomass yield calculated on oxygen (kg dw/kg O2) - Thiele modulus - Efficiency factor =1-(RO/R)3 (-) - Growth rate (m-1·s-1·kg1/3) - Dry weight per volume of pellet (kg·m-3)  相似文献   

12.
Changes in carp myosin ATPase induced by temperature acclimation   总被引:8,自引:0,他引:8  
Summary Myosins were isolated from dorsal ordinary muscles of carp acclimated to 10°C and 30°C for a minimum of 5 weeks and examined for their ATPase activities. Ca2+-ATPase activity was different between myosins from cold-and warm-acclimated carp, especially at KCl concentrations ranging from 0.1 to 0.2 M, when measured at pH 7.0. The highest activity was 0.32 mol Pi·min-1·mg-1 at 0.2 M KCl for cold-acclimated carp and 0.47 mol Pi·min-1·mg-1 at 0.1 M KCl for warm-acclimated fish. The pH-dependency of Ca2+-ATPase activity at 0.5 M KCl for both carp was, however, similar exhibiting two maxima around 0.3 mol Pi·min-1·mg-1 at pH 6 and 0.4 mol Pi·min-1·mg-1 at pH 9. K+(EDTA)-ATPase activity at pH 7.0 neither exhibited differences between both myosins. It increased with increasing KCl concentration showing the highest value of about 0.4 mol Pi·min-1·mg-1 at 0.6–0.7 M KCl. Actin-activated myosin Mg2+-ATPase activity was markedly different between cold-and warm-acclimated carp. The maximum initial velocity was 0.53 mol Pi·min-1·mg-1 myosin at pH 7.0 and 0.05 M KCl for cold-acclimated carp, which was 1.6 times as high as that for warm-acclimated carp. These differences were in good agreement with those obtained with myofibrillar Mg2+-ATPase activity between both carp. No differences were, however, observed in myosin affinity to actin. Differences in myosin properties between cold- and warm-acclimated carp were further evidenced by its thermal stability. The inactivation rate constant of myosin Ca2+-ATPase was 25·10-4·s-1 at 30°C and pH 7.0 for cold-acclimated carp, which was about 4 times as high as that for warm-acclimated carp. Light chain composition did not differ between both carp myosins. The differences in a primary structure of the heavy chain subunit was, however, clearly demonstrated between both myosins by peptide mapping.Abbreviations ATPase adenosine 5-triphosphatase - DTNB 5,5 dithio-bis-2-nitrobenzoic acid - DTT dithiothreitol - EGTA ethyleneglycol bis (-aminoethylether)-N,N,N,N-tetraacetic acid - K D inactivation rate constant - SDS sodium dodecyl sulfate - SDS-PAGE SDS-polyacrylamide gel electrophoresis  相似文献   

13.
Hemicellulose extracted from cell walls of suspension-cultured rose (Rosa Paul's Scarlet) cells was digested with cellulase from Trichoderma viride. The quantitatively major oligosaccharide products, a nonasaccharide and a heptasaccharide derived from xyloglucan, were purified by gel permeation chromatography. The nonasaccharide was found to inhibit the 2,4-dichlorophenoxy-acetic-acid-induced elongation of etiolated pea (Pisum sativum) stem segments. This confirms an earlier report (York et al., 1984, Plant Physiol. 75, 295–297). The inhibition of elongation by the nonasaccharide showed a maximum at around 10-9M with higher and lower concentrations being less effective. The heptasaccharide did not significantly inhibit elongation at 10-7–10-10M and also did not affect the inhibition caused by the nonasaccharide when co-incubated with the latter.Abbreviations 2,4-D 2,4-dichlorophenoxyacetic acid - XG xyloglucan - XG7 xyloglucan heptasaccharide (Glc4·Xyl3) - XG9 xyloglucan nonasaccharide (Glc4·Xyl3·Gal·Fuc)  相似文献   

14.
A comparative evaluation of the in vitro susceptibilities of 597 clinical yeast isolates to amphotericin B, fluconazole, and 5-fluorocytosine (5FC) was conducted. The broth macrodilution reference method of the National Committee for Clinical Laboratory Standards (NCCLS, M27-P) was adapted to the microdilution method. Microdilution endpoints for amphotericin B were scored as the lowest concentration in which a score of 0 (complete absence of growth) was observed and for 5FC and fluconazole as the lowest concentration in which a score of 2 (prominent decrease in turbidity; MIC-2) was observed compared to the growth control. The MIC values were read after 24 and 48 h incubation. A broad range of MIC values was observed with each antifungal agent. Amphotericin B was very active (MIC901.0 µg/ml) against all of the yeast isolates with the exception ofC. lusitaniae (MIC902.0 µg/ml). Fluconazole was most active againstC. parapsilosis (MIC90 of 1.0 µg/ml) and least active againstC. krusei (MIC90 of 32 µg/ml). 5FC was most active againstC. albicans, C. parapsilosis, C. tropicalis, andT. glabrata (MIC901.0 µg/ml) and was least active againstC. krusei andC. lusitaniae (MIC9016 µg/ml). These data indicate that the microdilution method, performed in accordance with M27-P, provides a means of testing larger numbers of yeast isolates against an array of antifungal agents and allows this to be accomplished in a reproducible and standardized manner. Given these results, it appears that the microdilution method may be a useful alternative to the macrodilution reference method for susceptibility testing of yeasts.  相似文献   

15.
We examined transepithelial transport of Ca2+ across the isolated opercular epithelium of the euryhaline killifish adapted to fresh water. The opercular epithelium, mounted in vitro with saline on the serosal side and fresh water (0.1 mmol·l–1 Ca2+) bathing the mucosal side, actively transported Ca2+ in the uptake direction; net flux averaged 20–30 nmol·cm–2·h–1. The rate of Ca2+ uptake varied linearly with the density of mitochondria-rich cells in the preparations. Ca2+ uptake was saturable, apparent K 1/2 of 0.348 mmol·l–1, indicative of a multistep transcellular pathway. Ca2+ uptake was inhibited partially by apically added 0.1 mmol·l–1 La3+ and 1.0 mmol·l–1 Mg2+. Addition of dibutyryl-cyclic adenosine monophosphate (0.5 mmol·l–1)+0.1 mmol·l–1 3-isobutyl-l-methylxanthine inhibited Ca2+ uptake by 54%, but epinephrine, clonidine and isoproterenol were without effect. Agents that increase intracellular Ca2+, thapsigargin (1.0 mol·l–1, serosal side), ionomycin (1.0 mol·l–1, serosal side) and the calmodulin blocker trifluoperazine (50 mol·l–1, mucosal side) all partially inhibited Ca2+ uptake. In contrast, apically added ionomycin increased mucosal to serosal unidirectional Ca2+ flux, indicating Ca2+ entry across the apical membrane is rate limiting in the transport. Verapamil (10–100 mol·l–1, mucosal side), a Ca2+ channel blocker, had no effect. Results are consistent with a model of Ca2+ uptake by mitochondria rich cells that involves passive Ca2+ entry across the apical membrane via verapamil-insensitive Ca2+ channels, intracellular complexing of Ca2+ by calmodulin and basolateral exit via an active transport process. Increases in intracellular Ca2+ invoke a downregulation of transcellular Ca2+ transport, implicating Ca2+ as a homeostatic mediator of its own transport.Abbreviations DASPEI 2-(4-dimethylaminostyryl)-N-ethylpyridinium iodide - db-cAMP dibutyryl-cyclic adenosine monophosphate - FW fresh water - G t transepithelial conductance - I sc short-circuit current - IBMX 3-isobutyl-1-methylxanthine - SW sea water - TFP trifluoperazine - V t transepithelial potential  相似文献   

16.
Gallbladder Na+ absorption is linked to gallstone formation in prairie dogs. We previously reported Na+/H+ exchanger (NHE1-3) expression in native gallbladder tissues. Here we report the functional characterization of NHE1, NHE2 and NHE3 in primary cultures of prairie dog gallbladder epithelial cells (GBECs). Immunohistochemical studies showed that GBECs grown to confluency are homogeneous epithelial cells of gastrointestinal origin. Electron microscopic analysis of GBECs demonstrated that the cells form polarized monolayers characterized by tight junctions and apical microvilli. GBECs grown on Snapwells exhibited polarity and developed transepithelial short-circuit current, Isc, (11.6 ± 0.5 µA · cm–2), potential differences, Vt (2.1 ± 0.2 mV), and resistance, Rt (169 ± 12 · cm2). NHE activity in GBECs assessed by measuring dimethylamiloride-inhibitable 22Na+ uptake under a H+ gradient was the same whether grown on permeable Snapwells or plastic wells. The basal rate of 22Na+ uptake was 21.4 ± 1.3 nmol · mg prot–1 · min–1, of which 9.5 ± 0.7 (~45%) was mediated through apically-restricted NHE. Selective inhibition with HOE-694 revealed that NHE1, NHE2 and NHE3 accounted for ~6%, ~66% and ~28% of GBECs total NHE activity, respectively. GBECs exhibited saturable NHE kinetics (Vmax 9.2 ± 0.3 nmol · mg prot–1 · min–1; Km 11.4 ± 1.4 mM Na+). Expression of NHE1, NHE2 and NHE3 mRNAs was confirmed by RT-PCR analysis. These results demonstrate that the primary cultures of GBECs exhibit Na+ transport characteristics similar to native gallbladder tissues, suggesting that these cells can be used as a tool for studying the mechanisms of gallbladder ion transport both under physiologic conditions and during gallstone formation.  相似文献   

17.
Endogenous and maximum respiration rates of nine purple sulfur bacterial strains were determined. Endogenous rates were below 10 nmol O2 · (mg protein · min)-1 for sulfur-free cells and 15–35 nmol O2 · (mg protein · min)-1 for cells containg intracellular sulfur globules. With sulfide as electron-donating substrate respiration rates were considerably higher than with thiosulfate. Maximum respiration rates of Thiocystis violacea 2711 and Thiorhodovibrio winogradskyi SSP1 (254.8 and 264.2 nmol O2 · (mg protein · min)-1, respectively) are similar to those of aerobic bacteria. Biphasic respiration curves were obtained for sulfur-free cells of Thiocystis violacea 2711 and Chromatium vinosum 2811. In Thiocystis violacea the rapid and incomplete oxidation of thiosulfate was five times faster than the oxidation of stored sulfur. A high affinity of the respiratoty system for oxygen (K m =0.3–0.9 M O2, V max=260 nmol O2 · (mg protein · min)-1 with sulfide as substrate, K m =0.6–2.4 M O2, V max=14–40 nmol O2 · (mg protein · min)-1 with thiosulfate as substrate), for sulfide (K m =0.47 M, V max=650 nmol H2S · (mg protein × min)-1, and for thiosulfate (K m =5–6 M, V max =24–72 nmol S2O 3 2- · (mg protein · min)-1 was obtained for different strains. Respiration of Thiocystis violacea was inhibited by very low concentrations of NaCN (K i =1.7 M) while CO concentrations of up to 300 M were not inhibitory. The capacity for chemotrophic growth of six species was studied in continuous culture at oxygen concentrations of 11 to 67 M. Thiocystis violacea 2711, Amoebobacter roseus 6611, Thiocapsa roseopersicina 6311 and Thiorhodovibrio winogradskyi SSP1 were able to grow chemotrophically with thiosulfate/acetate or sulfide/acetate. Chromatium vinosum 2811 and Amoebobacter purpureus ML1 failed to grow under these conditions. During shift from phototrophic to chemotrophic conditions intracellular sulfur and carbohydrate accumulated transiently inside the cells. During chemotrophic growth bacteriochlorophyll a was below the detection limit.  相似文献   

18.
Summary Regulation of the paracellular pathway in rabbit distal colon by the hormone aldosterone was investigated in vitro in Ussing chambers by means of transepithelial and microelectrode techniques. To evaluate the cellular and paracellular resistances an equivalent circuit analysis was used. For the analysis the apical membrane resistance was altered using the antibiotic nystatin. Under control conditions two groups of epithelia were found, each clearly dependent on the light: dark regime. Low-transporting epithelia (LT) were observed in the morning and high-transporting epithelia (HT) in the afternoon. Na+ transport was about 3-fold higher in HT than in LT epithelia. Incubating epithelia of both groups with 0.1 mol·1-1 aldosterone on the serosal side nearly doubled in LT epithelia the short circuit current and transepithelial voltage but the transepithelial resistance was not influenced. Maximal values were reached after 4–5 h of aldosterone treatment. In HT epithelia due to the effect of aldosterone all three transepithelial parameters remained constant over time. Evaluation of the paracellular resistance revealed a significant increase after aldosterone stimulation in both epithelial groups. This increase suggests that tight junctions might have been regulated by aldosterone. The hormonal effect on electrolyte transport was also dependent on the physiological state of the rabbit colon. Since net Na+ absorption in distal colon is, in addition to transcellular absorption capacity, also dependent on the permeability of the paracellular pathway, the regulation of tight junctions by aldosterone may be a potent mechanism for improving Na+ absorption during hormone-stimulated ion transport.Abbreviations V t transepithelial potential difference (mV) - R t transepithelial resistance (·cm2) - G t transepithelial conductance (mS·cm-2) - Isc calculated short circuit current (A·cm-2) - V a apical membrane potential difference (mV) - V bl basolateral membrane potential difference (mV) - voltage divider ratio - R a apical membrane resistance (·cm2) - R bl basolateral membrane resistance (·cm2) - R c cellular resistance ( of apical and basolateral resistance) (·cm2) - R p resistance of the paracellular pathway (·cm2) - G a apical membrane conductance (mS·cm-2) - G bl basolateral membrane conductance (mS·cm-2) - G p paracellular conductance (mS·cm-2) - G t transepithelial conductance (mS·cm-2) - HT contr high transporting control epithelia - LT contr low transporting control epithelia - HT aldo aldosterone incubated high transporting epithelia - LT aldo aldosterone incubated low transporting epithelia  相似文献   

19.
The effect of temperature on the respiration rate of meiofauna   总被引:2,自引:0,他引:2  
R. Price  R. M. Warwick 《Oecologia》1980,44(2):145-148
Summary The effect of temperature on respiration rate has been established, using Cartesian divers, for the meiofaunal sabellid polychaeteManayunkia aestuarina, the free-living nematodeSphaerolaimus hirsutus and the harpacticoid copepodTachidius discipes from a mudflat in the Lynher estuary, Cornwall, U.K. Over the temperature range normally experienced in the field, i.e. 5–20° C the size-compensated respiration rate (R c) was related to the temperature (T) in °C by the equation Log10 R c=-0.635+0.0339T forManayunkia, Log10 R c=0.180+0.0069T forSphaerolaimus and Log10 R c=-0.428+0.0337T forTachidius, being equivalent toQ 10 values of 2.19, 1.17 and 2.17 respectively. In order to derive the temperature response forManayunkia a relationship was first established between respiration rate and body size: Log10 R=0.05+0.75 Log10 V whereR=respiration in nl·O2·ind-1·h-1 andV=body volume in nl.TheQ 10 values are compared with values for other species derived from the literature. From these limited data a dichotomy emerges: species with aQ 102 which apparently feed on diatoms and bacteria, the abundance of which are subject to large short term variability, and species withQ 101 apparently dependent on more stable food sources.  相似文献   

20.
During intracellular polarization of identified sensory neurons of the leech by square pulses of hyperpolarizing current electrical parameters of the cell membranes were determined: input resistance of the neuron Rn, time constant of the membrane , the ratio between conductance of the cell processes and conductance of the soma , the resistance of the soma membrane rs, the input resistance of the axon r a , capacitance of the membrane Cs, and resistivity of the soma membrane Rs. The results obtained by the study of various types of neurons were subjected to statistical analysis and compared with each other. Significant differences for neurons of N- and T-types were found only between the values of , Cs, and Rs (P<0.01). These parameters also had the lowest coefficients of variation. The surface area of the soma of the neurons, calculated from the capacitance of the membrane (the specific capacitance of the membrane was taken as 1 µF/cm2) was 7–10 times (N-neurons) or 4–6 times (T-neurons) greater than the surface area of a sphere of the same diameter. The resistivity of the soma membrane Rs was 35.00 k·cm2 for cells of the N-type and 19.50 k·cm2 for T-neurons. The reasons for the relative stability of this parameter compared with the input resistance of the cell (coefficient of variation 22–7 and 53–31% respectively) are discussed. The possible effects of electrical characteristics on the properties of repeated discharges in neurons of different types also are discussed.A. A. Zhdanov Leningrad State University. Translated from Neirofiziologiya, Vol.7, No.3, pp.295–301, May–June, 1975.  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