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 895 毫秒
1.
We have compared the movement of a series of fluorescent tracers of increasing molecular weight injected into the cytoplasm in the epidermal cells of leaves of Egeria densa Planch. In general, the tracers showed major movement into three cellular compartments: first, to the cytoplasm of adjacent cells; secondly, from the cytoplasm, to the vacuole (irreversible); and thirdly, from the cytoplasm to the nucleus (reversible). No visible accumulation in chloroplasts or mitochondria, or loss across the plasmalemma was observed. No evidence for metabolic breakdown was found in extracts from injected leaves. The time course of accumulation of the dye in the three major compartments (cytoplasm, nucleus, vacuole) was monitored using fluorescence microscopy. The rate measurements and the quantified geometry of the cells were used to generate a model of compartmentation during intercellular transport. Permeability coefficients were calculated and related to the molecular sizes of the tracers. The coefficients for the tonoplast and nuclear envelope were independent of the molecular sizes of the tracers, and were in the range 2.4·10–6–4.1· 10–6 cm·s–1 for the tonoplast, and 2.6·10–5-9.4.10–5 cm· s–1 for the nuclear envelope. For intercellular movement, permeabilities were strongly dependent on molecular size, and ranged from 1.1·10–4 cm·s–1 for 6-carboxyfluorescein (376 daltons (Da)) to 9·10–9 cm·s–1 for fluorescein leucyldiglutamylleucine (874 Da). Thus, the differences in cell-to-cell movement of these tracers are based upon their differing ability to cross the intercellular walls, not upon differences in their intracellular compartmentation.Abbreviations 6COOHF 6-Carboxyfluorescein - Da daltons - DAPI 4,6-diamidino-2-phenylindole - F fluorescein-isothiocyanate isomer I - FGlu fluorescein glutamic acid - F(Glu)2 fluorescein glutamyl-glutamic acid - F(Gly)6 fluorescein hexaglycine - FLGGL fluorescein leucyl diglutamyl-leucine This work was supported by the Australian Research Grant Scheme. The assistance of Professor B.E.S. Gunning (Australian National University, Canberra, ACT) in providing facilities for making the photometer measurements is gratefully acknowledged.  相似文献   

2.
The hydraulic conductivities of excised whole root systems of wheat (Triticum aestivum L. cv. Atou) and of single excised roots of wheat and maize (Zea mays L. cv. Passat) were measured using an osmotically induced back-flow technique. Ninety minutes after excision the values for single excised roots ranged from 1.6·10-8 to 5.5·10-8 m·s-1·MPa-1 in wheat and from 0.9·10-8 to 4.8·10-8 m·s-1·MPa-1 in maize. The main source of variation was a decrease in the value as root length increased. The hydraulic conductivities of whole root systems, but not of single excised roots, were smaller 15 h after excision. This was not caused by occlusion of the xylem at the cut end of the coleoptile. The hydraulic conductivities of epidermal, cortical and endodermal cells were measured using a pressure probe. Epidermal and cortical cells of both wheat and maize roots gave mean values of 1.2·10-7 m·s-1·MPa-1 but in endodermal cells (measured only in wheat) the mean value was 0.5·10-7 m·s-1·MPa-1. The cellular hydraulic conductivities were used to calculate the root hydraulic conductivities expected if water flow across the root was via transcellular (vacuole-to-vacuole), apoplasmic or symplasmic pathways. The results indicate that, in freshly excised roots, the bulk of water flow is unlikely to be via the transcellular pathway. This is in contrast to our previous conclusion (H. Jones, A.D. Tomos, R.A. Leigh and R.G. Wyn Jones 1983, Planta 158, 230–236) which was based on results obtained with whole root systems of wheat measured 14–15 h after excision and which probably gave artefactually low values for root hydraulic conductivity. It is now concluded that, near the root tip, water flow could be through a symplasmic pathway in which the only substantial resistances to water flow are provided by the outer epidermal and the inner endodermal plasma membranes. Further from the tip, the measured hydraulic conductivities of the roots are consistent with flow either through the symplasmic or apoplasmic pathways.Symbols L p, cell cell hydraulic conductivity - L p, root root hydraulic conductivity - L p, root calculated root hydraulic conductivity - root reflection coefficient  相似文献   

3.
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   

4.
In C4 grasses belonging to the NADP-malic enzyme-type subgroup, malate is considered to be the predominant C4 acid metabolized during C4 photosynthesis, and the bundle sheath cell chloroplasts contain very little photosystem-II (PSII) activity. The present studies showed that Flaveria bidentis (L.), an NADP-malic enzyme-type C4 dicotyledon, had substantial PSII activity in bundle sheath cells and that malate and aspartate apparently contributed about equally to the transfer of CO2 to bundle sheath cells. Preparations of bundle sheath cells and chloroplasts isolated from these cells evolved O2 at rates between 1.5 and 2 mol · min–1 · mg–1 chlorophyll (Chl) in the light in response to adding either 3-phosphoglycerate plus HCO 3 or aspartate plus 2-oxoglutarate. Rates of more than 2 mol O2 · min–1 · mg–1 Chl were recorded for cells provided with both sets of these substrates. With bundle sheath cell preparations the maximum rates of light-dependent CO2 fixation and malate decarboxylation to pyruvate recorded were about 1.7 mol · min–1 · mg–1 Chl. Compared with NADP-malic enzyme-type grass species, F. bidentis bundle sheath cells contained much higher activities of NADP-malate dehydrogenase and of aspartate and alanine aminotransferases. Time-course and pulse-chase studies following the kinetics of radiolabelling of the C-4 carboxyl of C4 acids from 14CO2 indicated that the photosynthetically active pool of malate was about twice the size of the aspartate pool. However, there was strong evidence for a rapid flux of carbon through both these pools. Possible routes of aspartate metabolism and the relationship between this metabolism and PSII activity in bundle sheath cells are considered.Abbreviations DHAP dihydroxyacetone phosphate - NADP-ME(-type) NADP-malic enzyme (type) - NADP-MDH NADP-malate dehydrogenase - OAA oxaloacetic acid - 2-OG 2-oxoglutarate - PEP phosphoenolpyruvate - PGA 3-phosphoglycerate - Pi orthophosphate - Ru5P ribulose 5-phosphate  相似文献   

5.
The stroke volume of the left ventricle (SV) was calculated from noninvasive recordings of the arterial pressure using a finger photoplethysmograph and compared to the values obtained by pulsed Doppler echocardiography (PDE). A group of 19 healthy men and 12 women [mean ages: 20.8 (SD 1.6) and 22.2 (SD 1.6) years respectively] were studied at rest in the supine position. The ratio of the area below the ejection phase of the arterial pressure wave (A s) to SV, as obtained by PDE, yielded a calibration factor dimensionally equal to the hydraulic impedance of the system (Z ao =A s ·SV –1). TheZ ao amounted on average to 0.062 (SD 0.018) mmHg · s · cm–3 for the men and to 0.104 (SD 0.024) mmHg · s · cm–3 for the women. TheZ ao was also estimated from the equation:Z ao = a · (d + b ·HR + c ·PP + e ·MAP)–1, whereHR was the heart rate,PP the pulse pressure,MAP the mean arterial pressure and the coefficients of the equation were obtained by an iterating statistical package. The value ofZ ao thus obtained allowed the calculation of SV from measurements derived from the photoplethysmograph only. The mean percentage error between the SV thus obtained and those experimentally determined by PDE amounted to 14.8 and 15.6 for the men and the women, respectively. The error of the estimate was reduced to 12.3 and to 11.1, respectively, if the factorZ ao, experimentally obtained from a given heart beat, was subsequently applied to other beats to obtain SV from theA s measurement in the same subject.  相似文献   

6.
Production of hydrogen peroxide has been found in Ulva rigida (Chlorophyta). The formation of H2O2 was light dependent with a production of 1.2 mol·g FW–1·h–1 in sea water (pH 8.2) at an irradiance of 700 mol photons m–2·s–1. The excretion was also pH dependent: in pH 6.5 the production was not detectable (< 5 nmol·g FW–1·h–1) but at pH 9.0 the production was 5.0 mol·g FW–1·h–1. The production of H2O2 was totally inhibited by 3-(3,4-dichlorophenyl)-1,1 dimethylurea (DCMU). The ability of U. rigida growing in tanks (7501) under a natural light regime to excrete H2O2 was checked and found to be seven times higher at 08.00 hours than other times of the day. The H2O2 concentration in the cultivation tank (density: 2 g FW·l–1) reached the highest value (3 M) at 11.00 hours. Photosynthesis was not influenced by H2O2 formation. The H2O2 is suggested to come from the Mehler reaction (pseudocyclic photophosphorylation). With an oxygen evolution of 120 mmol·g FW–1·h–1 at pH 8.2 and 90 mmol·g FW–1·h–1 at pH 9.0, 0.5% and 2.7% of the electrons were used for extracellular H2O2 production. The H2O2 production is sufficiently high to be of physiological and ecological significance, and is suggested to be a part of the defence against epi and endophytes.Abbreviations ACL artificial, continuous light - DCMU 3-(3,4-dichlorophenyl)-1,1-dimethylurea - GNL greenhouse - LDC Luminol-dependent chemiluminescence - SOD Superoxide dismutase This investigation was supported by SAREC (Swedish Agency for Research Cooperation with Developing Countries), Hierta-Retzius Foundation, Marianne and Marcus Wallenberg Foundation, the Swedish Environmental Protection Board, and CICYT Spain.  相似文献   

7.
Using primary cultures of gill pavement cells from freshwater rainbow trout, a method is described for achieving confluent monolayers of the cells on glass coverslips. A continuous record of intracellular pH was obtained by loading the cells with the pH-sensitive flourescent dye 2,7-bis(2-carboxyethyl)-5(6)-carboxyfluorescein and mounting the coverslips in the flowthrough cuvette of a spectrofluorimeter. Experiments were performed in HEPES-buffered media nominally free of HCO3. Resting intracellular pH (7.43 at extracellular pH=7.70) was insensitive to the removal of Cl or the application of 4-acetamido-4-isothiocyanatostilbene-2,2-disulfonic acid (0.1 mmol·l–1), but fell by about 0.3 units when Na+ was removed or in the presence of amiloride (0.2 mmol·l–1). Exposure to elevated ammonia (ammonia prepulse; 30 mmol·l–1 as NH4Cl for 6–9 min) produced an increase in intracellular pH (to about 8.1) followed by a slow decay, and washout of the pulse caused intracellular pH to fall to about 6.5. Intracellular non-HCO 3 buffer capacity was about 13.4 slykes. Rapid recovery of intracellular pH from intracellular acidosis induced by ammonia prepulse was inhibited more than 80% in Na+-free conditions or in the presence of amiloride (0.2 mmol·l–1). Neither bafilomycin A1 (3 mol·l–1) nor Cl removal altered the intracellular pH recovery rate. The K m for Na+ of the intracellular pH recovery mechanism was 8.3 mmol·l–1, and the rate constant at V max was 0.008·s–1 (equivalent to 5.60 mmol H+·l–1 cell water·min–1), which was achieved at external Na+ levels from 25 to 140 mmol·l–1. We conclude that intracellular pH in cultured gill pavement cells in HEPES-buffered, HCO 3 -free media, both at rest and during acidosis, is regulated by a Na+/H+ antiport and not by anion-dependent mechanisms or a vacuolar H+-ATPase.Abbreviations BCECF 2,7-bis(2-carboxyethyl)-5(6)-carboxy-fluorescein - BCECF/AM 2,7-bis(2-carboxyethyl)-5(6)-carboxy-fluorescein, acetoxymethylester - Cholin-Cl choline chloride - DMSO dimethyl sulfoxide - EDTA ethylene diamine tetra-acetic acid - FBS foetal bovine serum - H + -ATPase Proton-dependent adenosine triphosphatase - HEPES N-[2-hydroxyethyl]piperazine-N[2-ethanesulfonic acid] - pH i intracellular pH - pH e extracellular pH - PBS phosphate-buffered saline - SITS 4-acetamido-4-isothiocyanatostilbene-2,2-disulfonic acid  相似文献   

8.
Zusammenfassung Die Reaktionskinetik strahleninduzierter freier Radikale des Cholesterins wurde in flüssiger Phase bei Raumtemperatur mittels ESR-Spektroskopie untersucht. Mit Hilfe eines geeigneten photochemischen Initiationssystems ließen sich in Cyclohexanlösung unter UV-Bestrahlung (235 nm265 nm) genau dieselben freien Radikale des Cholesterins darstellen, die schon früher [9, 7] in röntgenbestrahltem Cholesterinpulver beobachtet worden waren. Bei ausreichendem O2-Partialdruck (3·104Torr) über der Probenlösung trat das ESR-Spektrum eines Peroxyradikals auf, das mittels der Analyse seiner Reaktionsprodukte (7-Hydroxy-Cholesterin und 7-Keto-Cholesterin) mit dem Cholesteryl-7-peroxyradikal identifiziert wurde. Die Kinetik sowohl der Bildung als auch des Zerfalls des Radikals entsprachen einer Reaktion von 2. Ordnung. Die Geschwindigkeitskonstante für den bimolekularen Zerfall, eine Disproportionierung in Alkohol und Keton unter Abgabe eines Moleküls O2, wurde bei Raumtemperatur zuk 2=(1,8 –0,6 +0,9 )·106 sec–1M–1·l bestimmt. Ferner wurde gezeigt, daß das Cholesteryl-7-peroxyradikal aus dem freien Radikal Cholesteryl-7 durch Anlagerung eines Moleküls O2 entsteht. Für die Geschwindigkeitskonstante dieser Reaktion ergab sich eine untere Schranke vonk 1=0,40·1010 sec–1M–1·l.
Electron spin resonance investigations on radiation-induced free radicals of cholesterol in liquid phase
Summary The reaction kinetics of radiation-induced free radicals of cholesterol was studied in liquid phase at room temperature by means of e.s.r. spectroscopy on a solution of cholesterol in cyclohexane. Using a convenient photochemical initiation system, just those free radicals of cholesterol could be generated by the filtered u.v. radiation from a Xe high pressure lamp (235 nm265 nm) as were observed already a decade ago by Gordy [9] and by Ehrenberg, Löfroth [7] in X-irradiated cholesterol powder. At sufficiently high O2-pressures (3·10–4 Torr) over the sample solution a peroxy radical e.s.r. spectrum arose during u.v. irradiation which was identified by product analysis (7-hydroxy-cholesterol and 7-keto-cholesterol) to be dueto a cholesteryl-7-peroxyradical. The radical'sgeneration and decay kinetics was governed by a second order reaction. The velocity constant for bimolecular decay of the cholesteryl-7-peroxyradical was found to be k2=(1.8 –0,6 +0,9 )·106sec–1M–1·l at room temperature. Furthermore it could be shown that the cholesteryl-7-peroxyradical was built up by the addition of one molecule of O2 to a cholesteryl-7 free radical. For this reaction a value ofk 1=0.4·1010 sec–1 M–1·l was estimated as a lower limit of the velocity constant.


Die Arbeit stellt einen Auszug aus einer Dissertation an der Technischen Hochschule München dar.  相似文献   

9.
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   

10.
The ecology of Lake Nakuru (Kenya)   总被引:11,自引:0,他引:11  
E. Vareschi 《Oecologia》1982,55(1):81-101
Summary Abiotic factors, standing crop and photosynthetic production were studied in the equatorial alkaline-saline closed-basin Lake Nakuru (cond. 10,000–160,000 S). Meteorological conditions and abiotic factors offer suppositions for a high primary productivity: mean solar radiation is 450–550 kerg·cm-2·s-1, with little seasonal variation, regular winds circulate the lake every day and nutrient concentrations are usually high (>100 g P–PO4·l-1). Oxygen concentrations near sediments were <1 gO2·m-3 for at least 6 h·d-1 in 1972/73, resulting in a release of 45 mg P–PO4·m-2·d-1. Attenuation coefficients vary from 3.6–16.5 according to algal densities and mean depth from 0–400 cm. Algal biomass was 200 g·m-3 (d.w.) in 1972/73, due to a lasting Spirulina platensis bloom (98.5% of algal biomass). In 1974 algal biomass suddenly dropped to 50 g·m-3 (d.w.). Spirulina and several consumer organisms almost vanished, but coccoid cyanobacteria, Anabaenopsis and diatoms increased. Several causes for this change in ecosystem structure are discussed. The use of the light/dark bottle method to measure photosynthetic production in eutrophic alkaline lakes is discussed and relevant experiments were done. Oxygen tensions of 2–35 gO2·m-3 do not influence primary production rates. Net photosynthetic rates (mgO2·m-3·h-1; photosynthetic quotient=1.18) reached 12–17.7 in 1972/73 and 2–3 in 1974, but vertically integrated rates were only 1–1.4 in 1972/73 and 0.8 in 1974, and daily net photosynthetic rates (gO2·m-3·24 h-1) 3.5 in 1972/73 and 1 in 1974. 50% of areal rates were produced within the 10 most productive cm of the depth profile. The disproportion between high algal standing crops and relatively low production rates is due to self-shading of the algae, reducing the euphotic zone to 35 cm in 1972/73 and 77 cm in 1974. Efficiency of light utilization is 0.4–2%, varying with time of day and phytoplankton density. In situ efficiencies show an inverse relationship to light intensities. Photosynthetic rates of L. Nakuru remain within the range of other African lakes (0.1–3 gO2·m-2·h-1). The relation of O2 produced/Chl a of the euphotic zone is 50% lower then in tropical African freshwater lakes and conforms to lakes of temperate regions.  相似文献   

11.
Eicosapentaenoic (EPA) and docosahexaenoic (DHA) acid productivities from chemostat cultures of an isolate of Isochrysis galbana have been studied. The productivities reached in the interval of dilution rates between 0.0295 h–1 and 0.0355 h–1 were 1.5mg·1–1·h–1 for lipids, 300 g·1–1·h–1 for EPA and 130g1·1–1·h–1 for DHA. Furthermore, light attenuation by mutual shading, and agitation speed influences on growth and fatty acid composition were analysed. A model relating steady-state dilution rates to internal average light intensity has been proposed, the parameter values of which obtained by non-linear regression were: maximum specific growth rate (max)=0.0426 h–1; the affinity of cells to light (Ik) = 10.92 W·m–2; the exponent (n) = 5.13; regression coefficient (r 2)=0.9999. Correspondence to: E. Molina Grima  相似文献   

12.
In a randomly selected sample of 88 men and 115 women, aged 23–27 years from Denmark, maximal oxygen uptake ( O2max), maximal voluntary isometric contraction (MVC) in four muscle groups and physical activity were studied. The O2max was 48.0 ml · min–1 kg–1 and 39.6 ml · min–1 · kg–1 for the men and the women, respectively. The MVC was 10% lower than in a comparable group of Danes of the same age and height studied 35 years ago. Only in men was sports activity directly related to O2max (ml · min–1 · kg–1; r=0.31, P<0.01). The MVC of the knee extensors was related to O2max in the men (r=0.31, P<0.01), but there was no relationship between the other measurements of MVC and O2max. In the women O2max (ml · min–1 · kg–1) was only related to body size, i.e. body mass index, percentage body fat and body mass [(r= –0.47, –0.48 (both P<0.001) and –0.34. (P<0.01), respectively)]. There were differences in O2max in the men, according to education and occupation. Blue collar workers and subjects attending vocational or trade schools in 1983 had lower O2max and more of them were physically inactive. In the women differences were also found, but there was no clear pattern among the groups. More of the women participated regularly in sports activity, but more of the men were very active compared to the women.  相似文献   

13.
The diving and thermoregulatory metabolic rates of two species of diving seabrid, common (Uria aalge) and thick-billed murres (U. lomvia), were studied in the laboratory. Post-absorptive resting metabolic rates were similar in both species, averaging 7.8 W·kg-1, and were not different in air or water (15–20°C). These values were 1.5–2 times higher than values predicted from published allometric equations. Feeding led to increases of 36 and 49%, diving caused increases of 82 and 140%, and preening led to increases of 107 and 196% above measured resting metabolic rates in common and thick-billed murres, respectively. Metabolic rates of both species increased linearly with decreasing water temperature; lower critical temperature was 15°C in common murres and 16°C in thick-billed murres. Conductance (assuming a constant body temperature) did not change with decreasing temperature, and was calculated at 3.59 W·m-2·oC-1 and 4.68 W·m-2·oC-1 in common and thick-billed murres, respectively. Murres spend a considerable amount of time in cold water which poses a significant thermal challenge to these relatively small seabirds. If thermal conductance does not change with decreasing water temperature, murres most likely rely upon increasing metabolism to maintain body temperature. The birds probably employ activities such as preening, diving, or food-induced thermogenesis to meet this challenge.Abbreviations ADL aerobic dive limit - BMR basal metabolic rate - FIT food-induced thermogenesis - MHP metabolic heat production - MR metabolic rate - PARR post-absorption resting rate - RMR resting metabolic rate - RQ respiratory quotient - SA surface area - STPD standard temperature and pressure (25°C, 1 ATM) - T a ambient temperature - T b body temperature - T IC Iower critical temperatiure - TC thermal conductance - V oxygen consumption rate - W body mass  相似文献   

14.
Summary Median volumes ofin vitro coelomocyte populations fromGlycera dibranchiata rapidly change in response to external differences of osmotic pressure (Fig. 3). Fifty per cent haemolysis occurred in just under 30% sea water, 285 mOSm·kg–1. Hydraulic conductivities (Lp=0.92 to 2.78×107 cm×s–1·atm–1) calculated from rates of osmotic swelling were similar to values for sea urchin eggs and squid axons. Coelomocytes show a slower partial return to their original volumes in hypotonic but not hypertonic media. This asymmetry is reflected in Ponder's R values of 0.787 and 0.987, respectively, determined after this regulatory phase is complete (Figs. 4 and 5). Evidence for an irreversible stress dependent leakage of osmotically active solutes when the coelomocytes are removed from the animal and diluted with sea water is presented.  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.
The stationary radial volume flows across maize (Zea mays L.) root segments without steles (sleeves) were measured under isobaric conditions. The driving force of the volume flow is an osmotic difference between the internal and external compartment of the root preparations. It is generated by differences in the concentrations of sucrose, raffinose or polyethylene glycol. The flows are linear functions of the corresponding osmotic differences ( ) up to osmotic values which cause plasmolysis. The straight lines obtained pass through the origin. No asymmetry of the osmotic barrier could be detected within the range of driving forces applied ( =±0.5 MPa), corresponding to volume-flow densities of jv, s=±7·10–8 m·s–1. Using the literature values for the reflection coefficients of sucrose and polyethylene glycol in intact roots (E. Steudle et al. (1987) Plant Physiol.84, 1220–1234), values for the sleeve hydraulic conductivity of about 1·10–7 m·s–1 MPa–1 were calculated. They are of the same order of magnitude as those reported in the literature for the hydraulic conductivity of intact root segments when hydrostatic pressure is applied.Abbreviations and symbols a s outer surface of sleeve segment - c concentration of osmotically active solute - j v, s radial volume flow density across sleeve segment - Lps hydraulic conductivity of sleeves - Lpr hydraulic conductivity of intact roots - N thickness of Nernst diffusion layer - reflection coefficient of root for solute - osmotic value of bulk phase - osmotic coefficient  相似文献   

17.
To determine the effect of endogenous opioids on catecholamine response during intense exercise [80% maximal oxygen uptake ( O2max)], nine fit men [mean (SE) ( O2max, 63.9 (1.7) ml · kg–1 · min–1; age 27.6 (1.6) years] were studied during two treadmill exercise trials. A double-blind experimental design was used with subjects undertaking the two exercise trials in counterbalanced order. Exercise trials were 20 min in duration and were conducted 7 days apart. One exercise trial was undertaken following administration of naloxone (N; 1.2 mmol · l–1; 3 ml) and the other after receiving a placebo (P; 0.9% saline; 3 ml). Prior to each experimental trial a flexible catheter was placed into an antecubital vein and baseline blood samples were collected. Immediately afterwards, each subject received bolus injection of either N or P. Blood samples were also collected after 20 min of continuous exercise while running. Epinephrine and norepinephrine were higher (P < 0.05) in the N than P exercise trial with mean (SE) values of 1679 (196) versus 1196 (155) pmol · l–1 and 24 (2.2) versus 20 (1.7) nmol · · l–1 respectively. Glucose and lactate were higher (P < 0.05) in the N than P exercise trial with values of 7 (0.37) versus 5.9 (0.31) mmol · l–1 and 6.9 (1.1) versus 5.3 (0.9) mmol · l–1 respectively. These data suggest an opioid inhibition in the release of catecholamines during intense exercise.  相似文献   

18.
Summary Instantaneous oxygen consumption, muscle potential frequency, thoracic and ambient temperature were simultaneously measured during heating in individual workers and drones of honey bees. Relationships between these parameters and effects of thoracic temperature on power input and temperature elevation were studied. Oxygen consumption increased above basal levels only when flight muscles became active. Increasing muscle potential frequencies correlated with elevated oxygen consumption and raised thoracic temperature. The difference between thoracic and ambient temperature and oxygen consumption were linearly related. Oxygen consumption per muscle potential (l O2 · g –1 thorax · MP–1) was two-fold higher in drones than in workers. However, oxygen consumption for heating the thorax (l O2 · g –1 thorax · (Tth-Ta) · °C–1) was nearly the same in workers and drones. Thoracic temperature affected the amount of oxygen consumed per muscle potential (R10=1.5). Achieved temperature elevation per 100 MP was more temperature sensitive in drones (R10=6–10) than in workers (R10=3.6). Q10 values for oxygen consumption were 3 in workers and 4.5–6 in drones. Muscle potential frequency decreased with a Q10=1.8 in workers and 2.7 in drones. Heating behaviour of workers and drones was different. Drones generated heat less continuously than workers, and showed greater interindividual variability in predilection to heat. However, the maximal difference between ambient and thoracic temperature observed was 22 °C in drones and 14 °C in workers, indicating greater potential for drones.Abbreviations DL dorsal-longitudinal muscle - DV dorsoventral muscle - MP muscle potential - T a ambient temperature - T th thoracic temperature  相似文献   

19.
Summary Unidirectional 22Na+ and 36Cl fluxes were determined in short-circuited, stripped rumen mucosa from sheep by using the Ussing chamber technique. In both CO2/HCO 3 -containing and CO2/HCO 3 -free solutions, replacement of gluconate by short-chain fatty acids (SCFA, 39 mM) significantly enhanced mucosal-toserosal Na+ absorption without affecting the Cl transport in the same direction. Short-chain fatty acid stimulation of Na+ transport was at least partly independent of Cl and could almost completely be abolished by 1 mM mucosal amiloride, while stimulation of Na+ transport was enhanced by lowering the mucosal pH from 7.3 to 6.5. Similar to the SCFA action, raising the PCO2 in the mucosal bathing solution led to an increase in the amiloride-sensitive mucosal-to-serosal Na+ flux. Along with its effect on sodium transport, raising the PCO2 also stimulated chloride transport. The results are best explained by a model in which undissociated SCFA and/or CO2 permeate the cell membrane and produce a raise in intracellular H+ concentration. This stimulates an apical Na+/H+ exchange, leading to increased Na+ transport. The stimulatory effect of CO2 on Cl transport is probably mediated by a Cl/HCO 3 exchange mechanism in the apical membrane. Binding of SCFA anions to that exchange as described for the rat distal colon (Binder and Mehta 1989) probably does not play a major role in the rumen.Abbreviations DIDS 4,4'-diisothiocyanatostilbene-2,2'-disulfonic acid - G t transepithelial conductance (mS·cm-2) - HSCFA undissociated short-chain fatty acids - J ms mucosal-to-serosal flux (Eq · cm-2 · h-1) - J net net flux (Eq · cm-2 · h-1) - J sm serosal-to-mucosal flux (Eq · cm-2 · h-1) - PD transepithelial potential difference (mV) - SCFA dissociated short-chain fatty acids - SCFA short-chain fatty acids  相似文献   

20.
Summary Rates of O2 uptake across isolated perfused skin of bullfrogs (Rana catesbeiana) were measured in relation to blood flow at three levels of ambient O2 tension: normoxia (O2 tension=152 torr), hypoxia (12% O2, 87 torr) and hyperoxia (42% O2, 306 torr). At bulk perfusion rates ranging from 3.4 to 10.1 l·cm-2·min-1, O2 uptake was positively correlated with hemoglobin delivery rate in both normoxia and hyperoxia, but was independent of delivery rate in hypoxia. Mean O2 uptake in normoxia was 3.8 nmol O2·cm-2·min-1 at a delivery rate of 9.8 nmol·cm-2·min-1 and 6.5 nmol O2·cm-2·min-1 at a delivery rate of 28.3 nmol·cm-2·min-1. At any given bulk perfusion rate, oxygen uptake averaged about 49% lower in hypoxia than in normoxia, decreasing in proportion to the reduction of O2 tension difference between medium and blood. In hyperoxia, O2 uptake did not increase proportionally with the difference in O2 tension between blood and medium, averaging only 50% higher at a 2.4-fold greater O2 tension difference. Cutaneous diffusing capacity for O2 averaged 0.041 nmol O2·cm-2·torr-1·min-1 during the first hour of perfusion in normoxia, and was not affected by reduction of ambient O2 tension. The results indicate that cutaneous O2 uptake in hypoxia is highly diffusion limited, and consequently, increases in cutaneous perfusion can not effectively compensate for reduction of ambient O2 tension. In hyperoxia, O2 uptake may be substantially perfusion limited because of reduced blood O2 capacitance at high O2 saturations.Abbreviations O2 capacitance - C Hb hemoglobin concentration - D diffusing capacity - PO2 medium-blood PO2 difference - Hb flow, hemoglobin delivery rate - Hepes N-[2-Hydroxyethyl]piperacine-N-[2 ethanesulfonic acid] - L diff extent of diffusion limitation - MO2 oxygen uptake rate - PO2 oxygen tension - S O2 saturation  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