首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 46 毫秒
1.
Trivalent chromium (Cr3+) is essential for animal and human health, whereas hexavalent Cr (CrO4 2−) is a potent carcinogen and extremely toxic to animals and humans. Thus, the accumulated Cr in food plants may represent potential health hazards to animals and humans if the element is accumulated in the hexavalent form or in high concentrations. This study was conducted to determine the extent to which various vegetable crops absorb and accumulate Cr3+ and CrO4 2− into roots and shoots and to ascertain the different chemical forms of Cr in these tissues. Two greenhouse hydroponic experiments were performed using a recirculating-nutrient culture technique that allowed all plants to be equally supplied with Cr at all times. In the first experiment, 1 mg L−1 Cr was supplied to 11 vegetable plant species as Cr3+ or CrO4 2−, and the accumulation of Cr in roots and shoots was compared. The crops tested included cabbage (Brassica oleracea L. var. capitata L.), cauliflower (Brassica oleracea L. var. botrytis L.), celery (Apium graveolens L. var. dulce (Mill.) Pers.), chive (Allium schoenoprasum L.), collard (Brassica oleracea L. var. acephala DC.), garden pea (Pisum sativum L.), kale (Brassica oleracea L. var. acephala DC.), lettuce (Lactuca sativa L.), onion (Allium cepa L.), spinach (Spinacia oleracea L.), and strawberry (Fragaria ×  ananassaDuch.). In the second experiment, X-ray absorption spectroscopy (XAS) analysis on Cr in plant tissues was performed in roots and shoots of various vegetable plants treated with CrO4 2− at either 2 mg Cr L−1 for 7 d or 10 mg Cr L−1 for 2, 4 or 7 d. The crops used in this experiment included beet (Beta vulgaris L. var. crassa (Alef.) J. Helm), broccoli (Brassica oleracea L. var. Italica Plenck), cantaloupe (Cucumis melo L. gp. Cantalupensis), cucumber (Cucumis sativus L.), lettuce, radish (Raphanus sativus L.), spinach, tomato (Lycopersicon lycopersicum (L.) Karsten), and turnip (Brassica rapa L. var. rapifera Bailey). The XAS speciation analysis indicates that CrO4 2− is converted in the root to Cr3+ by all plants tested. Translocation of both Cr forms from roots to shoots was extremely limited and accumulation of Cr by roots was 100-fold higher than that by shoots, regardless of the Cr species supplied. Highest Cr concentrations were detected in members of the Brassicaceae family such as cauliflower, kale, and cabbage. Based on our observations and previous findings by other researchers, a hypothesis for the differential accumulation and identical translocation patterns of the two Cr ions is proposed. Received: 27 February 1998 / Accepted: 2 April 1998  相似文献   

2.
The soluble acid invertase (β-D-fructofuranoside fructohydrolase, EC 3.2.1.26) from potato (Solanum tuberosum L. cv. Kennebec) tubers was located in the vacuoles. Although the functionality of this invertase in the vacuoles has been assumed, the activity of the enzyme has never been shown within isolated vacuoles. Vacuoles were prepared by gentle osmotic shock from free protoplasts obtained by enzymic digestion of tuber tissues. The mean volume of these vacuoles, (0.26 ± 0.05) × 10−2 μl, was estimated by optical microscopy. Sucrose, glucose and fructose concentrations were calculated to be 100 mM, 20 mM and 40 mM, respectively, in the vacuoles. Sucrose hydrolysis and the increase in glucose and fructose concentrations within the vacuoles were measured during vacuolar incubations. An almost identical pattern of sucrose hydrolysis by invertase was found by an in-vitro assay reproducing the vacuolar conditions. In view of the determinations of internal vacuolar pH (5.2), the possibility of spontaneous hydrolysis of sucrose was disregarded. Vacuoles were shown to be free from proteinaceous inhibitors, confirming the extravacuolar location of these inhibitors. The vacuolar hydrolytic pattern of sucrose confirms the regulatory role of the reaction products previously proposed for in-vitro assays. Received: 21 July 1997 / Accepted: 31 August 1997  相似文献   

3.
Brassinolide may control aquaporin activities in Arabidopsis thaliana   总被引:8,自引:0,他引:8  
 It is usually assumed that aquaporins present in the cellular membranes could be an important route in the control of water flux in plants, but evidence for this hypothesis is scarce. In this paper, we report measurements of the osmotic permeability (P os ) of protoplasts isolated from hypocotyls of wild-type and mutant Arabidopsis thaliana (L.) Heynh. Mutants were affected in their growth and exhibited different sensitivities to the phytohormone, brassinolide. For the two mutants studied (cpd: constitutive photomorphogenesis and dwarfism; bri1: brassinosteroid insensitive), hypocotyl length was correlated to P os for the protoplasts. Under experimental conditions where hypocotyl growth had ceased, restoration of root, hypocotyl and petiole growth by brassinolide was correlated with an increase in P os of the hypocotyl protoplasts. We consider that the increase in P os of the hypocotyl cells was needed because these cells were part of the transcellular water pathway of the plant. This is the first time, to our knowledge, that brassinolide has been shown to be involved in the modification of the water-transport properties of cell membranes. Our results also emphasize the importance of aquaporins and the transcellular pathway in water transport under normal growth conditions. Received: 15 January 2000 / Accepted: 18 May 2000  相似文献   

4.
The purposes of this study were to determine whether running economy (RE) is adversely affected following intense interval bouts of 10 × 400-m running, and whether there is an interaction effect between RE and recovery duration during the workouts. Twelve highly trained male endurance athletes [maximal oxygen consumption; O2 max =72.5 (4.3) ml·kg−1·min−1; mean (SD)] performed three interval running workouts of 10 × 400 m with a minimum of 4 days between runs. Recovery duration between the repetitions was randomly assigned at 60, 120 or 180 s. The velocity for each 400-m run was determined from a treadmill O2 max test. The average running velocity was 357.9 (9.0) m · min−1. Following the workout, the rating of perceived exertion (RPE) increased significantly (P < 0.01) as recovery duration between the 400-m repetitions decreased (14.4, 16.1, and 17.7 at 180s, 120s, and 60 s recovery, respectively). Prior to and following each workout, RE was measured at speeds of 200 and 268 m · min−1. Changes in RE from pre- to post-workout, as well as heart rate (HR) and respiratory exchange ratio (R) were similar for the three recovery conditions. When averaged across conditions, oxygen consumption (O2) increased significantly (P < 0.01) from pre- to post-test (from 38.5 to 40.5 ml · kg−1 · min−1 at 200 m · min−1, and from 53.1 to 54.5 ml · kg−1 · min−1 at 268 m · min−1, respectively). HR increased (from 124 to 138, and from 151 to 157 beats · min−1 respectively) and R decreased (from 0.90 to 0.78, and from 0.93 to 0.89, respectively) at 200 and 268 m · min−1, respectively (P < 0.01). This study showed that RE can be perturbed after a high-intensity interval workout and that the changes in O2, HR and R were independent of the recovery duration between the repetitions. Accepted: 23 June 1997  相似文献   

5.
Partitioning in a biphasic polymer system was used to isolate plasmalemma (PM) from roots and shoots of etiolated pea seedlings. The membrane preparations were used to assess the osmotic water permeability (P os) with the stopped-flow method. The Western-blot technique was employed to determine the membrane content of the PIP-family of aquaporins, and their activity was estimated by measuring the rate of osmotic vesicle shrinking in the presence of inhibitors, HgCl2 and AgNO3. Monobromobimane fluorescent dye was used to determine the quantity of sulfhydryl groups in cell membranes and follow the effect of SH-oxidizing (diamide) and SH-reducing (dithiothreitol and tributylphosphine) agents on P os of the root PM and oligomerization of aquaporins. The shoot PM was shown to combine high P os with low aquaporin content. In the root PM, P os was lower and the aquaporin content greatly exceeded that in the shoots. HgCl2 and AgNO3 did not decrease the rate of osmotic shrinking in root membrane vesicles, whereas considerably (by 40–50%) inhibited this index in the shoot membranes. Root and shoot PM preparations dramatically differed in their SH-group contents: the former exceeded the latter sixfold. When added to the homogenization medium, diamide and tributylphosphine affected the content of SH-groups and P os in the root PM. In the roots, diamide decreased the quantity of SH-groups almost twofold and increased P os fourfold, and the introduction of tributylphosphine produced a twofold increase in the quantity of SH-groups with only slight decrease in the P os. Immunological analysis of membranes isolated in the presence of diamide showed that the ratio between the monomer and dimer forms of aquaporins in two membrane preparations depended on the presence of dithiothreitol in the denaturing buffer apparently because dithiothreitol exposed and reduced disulfide bonds essential for monomer interactions and inaccessible for interaction with redox modifiers of SH-groups in the membrane. Because of their inaccessibility, these modifiers could not cause the changes of P os in the root PM produced by oxidation and reduction of SH-groups. This phenomenon is probably related to the change in the status of SH-groups in two cysteine residues at the N-end of the aquaporin loop C oriented outward into the apoplast.  相似文献   

6.
Oxygen consumption was measured in five Dermophis mexicanus and averaged (±SEM) 0.047 ± 0.004 ml O2 g−1 h−1. Carbon dioxide production averaged 0.053 ± 0.005 ml CO2 g−1 h−1 in the same five animals 1 week later. This metabolic rate is similar to metabolic rates of other Gymnophionans but lower than metabolic rates reported for Anurans and Urodeles. Total nitrogen excretion averaged 1.37 μmol N g−1 h−1 which is higher than that found for other amphibians. Of this, 82.5% (1.13 μmol N g−1 h−1) was in the form of urea while 17.5% (0.24 μmol N g−1 h−1) was in the form of NH3 + NH+ 4. Such ureotelism is typical of terrestrial amphibians like D. mexicanus. Osmotic water flux averaged 0.0193 ml g−1 h−1 in control (sham injected) animals and was not significantly altered by injection of either arginine vasotocin or mesotocin. This osmotic flux is similar to osmotic fluxes found for other terrestrial amphibians. The combined data suggest that metabolism in D. mexicanus is, like most other Gymnophionans, lower than other amphibians. The high rates of nitrogen (especially urea) excretion suggests that this fossorial animal accumulates urea like other burrowing amphibians. Accepted: 27 June 2000  相似文献   

7.
Transepithelial water permeability was measured in LLC-PK1 cells stably transfected with aquaporins (AQPs): AQP1, AQP2, and a chimera of AQP1 and AQP2 containing 41 amino acids of the C-terminus of AQP2. Transepithelial water fluxes (Jw) were not previously reported in cells transfected with aquaporins. Jw were now recorded each minute using a specially developed experimental device. A significant increase in Posm after forskolin (FK) plus vasopressin (VP) was found in AQP2 transfected cells (39.9 ± 8.2 vs. 12.5 ± 3.3 cm · sec−1· 10−3), but not in cells transfected with AQP1 (15.3 ± 3.6 vs. 13.4 ± 3.6 cm · sec−1· 10−3). In the case of the AQP1/2 cells (chimera) the FK plus VP induced Posm was smaller than in AQP2 cells but significantly higher than in mock cells at rest (18.1 ± 4.8 vs. 6.7 ± 1.0 cm · sec−1· 10−3). The increases in Posm values were not paralleled by increases in 14C-Mannitol permeability. HgCl2 inhibited the hydrosmotic response to FK plus VP in AQP2 transfected epithelia. Results were comparable to those observed, in parallel experiments, in a native ADH-sensitive water channel containing epithelial barrier (the toad urinary bladder). Electron microscopy showed confluent LLC-PK1 cells with microvilli at the mucosal border. The presence of spherical or elongated intracellular vacuoles was observed in AQP2 transfected cells, specially after FK plus VP stimulus and under an osmotic gradient. These results demonstrate regulated transepithelial water permeability in epithelial cells transfected with AQP2. Received: 24 June 1997/Revised: 16 September 1997  相似文献   

8.
Intra- and transcellular water movements in plants are regulated by the water permeability of the plasma membrane (PM) and vacuolar membrane (VM) in plant cells. In the present study, we investigated the osmotic water permeability of both PM (P f1) and VM (P f2), as well as the bulk osmotic water permeability of a protoplast (P f(bulk)) isolated from radish (Raphanus sativus) roots. The values of P f(bulk) and P f2 were determined from the swelling/shrinking rate of protoplasts and isolated vacuoles under hypo- or hypertonic conditions. In order to minimize the effect of unstirred layer, we monitored dropping or rising protoplasts (vacuoles) in sorbitol solutions as they swelled or shrunk. P f1 was calculated from P f(bulk) and P f2 by using the ‘three-compartment model’, which describes the theoretical relationship between P f1, P f2 and P f(bulk) (Kuwagata and Murai-Hatano in J Plant Res, 2007). The time-dependent changes in the volume of protoplasts and isolated vacuoles fitted well to the theoretical curves, and solute permeation of PM and VM was able to be neglected for measuring the osmotic water permeability. High osmotic water permeability of more than 500 μm s−1, indicating high activity of aquaporins (water channels), was observed in both PM and VM in radish root cells. This method has the advantage that P f1 and P f2 can be measured accurately in individual higher plant cells. Electronic supplementary material Supplementary material is available in the online version of this article at and is accessible for authorized users. It includes four appendices, four tables and two figures. Mari Murai-Hatano and Tsuneo Kuwagata contributed equally to the paper. An erratum to this article is available at .  相似文献   

9.
The aim of the present study was to examine the physiological and mechanical factors which may be concerned in the increase in energy cost during running in a fatigued state. A group of 15 trained triathletes ran on a treadmill at velocities corresponding to their personal records over 3000m [mean 4.53 (SD 0.28) m · s−1] until they felt exhausted. The energy cost of running (C R) was quantified from the net O2 uptake and the elevation of blood lactate concentration. Gas exchange was measured over 1 min firstly during the 3rd–4th min and secondly during the last minute of the run. Blood samples were collected before and after the completion of the run. Mechanical changes of the centre of mass were quantified using a kinematic arm. A significant mean increase [6.9 (SD 3.5)%, P < 0.001] in C R from a mean of 4.4 (SD 0.4) J · kg−1 · m−1 to a mean of 4.7 (SD 0.4) J · kg−1 · m−1 was observed. The increase in the O2 demand of the respiratory muscles estimated from the increase in ventilation accounted for a considerable proportion [mean 25.2 (SD 10.4)%] of the increase in CR. A mean increase [17.0 (SD 26.0)%, P < 0.05] in the mechanical cost (C M) from a mean of 2.36 (SD 0.23) J · kg−1 · m−1 to a mean of 2.74 (SD 0.55) J · kg−1 · m−1 was also noted. A significant correlation was found between C R and C M in the non-fatigued state (r = 0.68, P < 0.01), but not in the fatigued state (r = 0.25, NS). Furthermore, no correlations were found between the changes (from non-fatigued to fatigued state) in C R and the changes in C M suggesting that the increase in C R is not solely dependent on the external work done per unit of distance. Since step frequency decreased slightly in the fatigued state, the internal work would have tended to decrease slightly which would not be compatible with an increase in C R. A stepwise regressions showed that the changes in C R were linked (r = 0.77, P < 0.01) to the changes in the variability of step frequency and in the variability of potential cost suggesting that a large proportion of the increase in C R was due to an increase in the step variability. The underlying mechanisms of the relationship between C R and step variability remains unclear. Accepted: 15 September 1997  相似文献   

10.
Changes of vacuolar pH in hair cells of young rice (Oryza sativa L.) and maize (Zea mays L.) roots were measured after ammonia application at various levels of external pH. After loading the pH-sensitive, fluorescent dye Oregon green 488 carboxylic acid 6-isomer into the vacuoles of root hairs, ratiometric pH data of high statistical significance were obtained from root hair populations comprising hundreds of cells. The pH of the vacuole at external pH 5.0 was 5.32 ± 0.08 (±SD, n= 15) and 5.41 ± 0.13 (±SD, n= 15) in rice and maize, respectively. A moderate external ammonia concentration of 2 mM led to vacuolar alkalisation at both, low (pH 5.0) and high (pH 7.0–9.0) external pH, presumably due to NH3 permeation into the vacuole. With increasing external pH, ammonia application did not cumulatively increase vacuolar pH. In rice, the increase in vacuolar pH ranged from 0.1–0.8 pH units; in maize a more constant increase of 0.5 pH units was observed. The vacuolar pH increase was efficiently depressed in rice (especially at high external pH), but not in maize. Inhibition of the tonoplast H+-ATPase by concanamycin A raised vacuolar pH and increased the ammonia-elicited vacuolar alkalisation in both species, proving that vacuolar H+-ATPase activity counters the ammonia-elicited alkalisation effect. However, even under conditions of vacuolar H+-ATPase inhibition, rice was still able to restore an ammonia-elicited pH increase. High vacuolar pH levels as found in maize under conditions of high NH3 influx may derive from inefficient cytosolic ammonia assimilation and tonoplast proton pumping. Thus, in maize, prolonged reduction of the proton gradient between the cytosol and the vacuole may play an important role in NH3 toxicity. Received: 12 September 1997 / Accepted: 19 January 1998  相似文献   

11.
The mean effective water self-diffusion coefficient in maize root segments under the effect of aquaporin blocker (mercuric chloride, 0.1 mM) was measured using the spin-echo NMR method with pulsed magnetic field gradient within the temperature range from 10 to 35 °C. HgCl2 caused the reduction in water diffusion by 30 % as compared to the control samples. Temperature dependences of water self-diffusion coefficients showed two linear regions with different values of Q10 and activation energy, Ea. As the temperature reduced from 20 to 10 °C, Ea values calculated from the Arrhenius plots were close to those of bulk water (20 ± 3 kJ mol−1) and slightly changed for the sample pretreated HgCl2. Within the temperature range from 25 to 35 °C the slope of temperature dependences became steeper and Ea values were 31 ± 3 kJ mol−1 for the control and 40 ± 4 kJ mol−1 for the treated sample. In the vicinity of 20 °C, the temperature dependence of water diffusion via the mercury-sensitive water channels showed extreme value. In the region, the specific area of the mercury-sensitive aquaporins was 0.004 % of the total cell surface area. The data indicate that water transfer via aquaporins is sensitive to temperature, and the contributions of the transmembrane pathways (aquaporins, lipid bilayer) differ in different temperature ranges.  相似文献   

12.
The assumption that working on board ship is more strenuous than comparable work ashore was investigated in this study. Various physiological parameters (O2, CO2, E and HR) have been measured to determine the energy expenditure of subjects walking slowly on a moving platform (ship motion simulator). Twelve subjects (eight men and four women) walked either freely on the floor or on a treadmill at a speed of 1 m · s−1. Platform motion was either in a heave, pitch or roll mode. These three conditions were compared with a control condition in which the platform remained stationary. The results showed that during pitch and roll movements of the platform, the energy expenditure for the same walking task was about 30% higher than under the stationary control condition (3.6 J · kg−1 · m−1 vs 2.5 J · kg−1 · m−1, P < 0.05) for both walking on a treadmill and free walking. The heart rate data supported the higher energy expenditure results with an elevation of the heart rate (112 beats · min−1 vs 103 beats · min−1, P < 0.05). The heave condition did not differ significantly from the stationary control condition. Pitch and roll were not significantly different from each other. During all experimental conditions free walking resulted in a higher energy cost of walking than treadmill walking (3.5 J · kg−1 · m−1 vs 2.7 J · kg−1 · m−1, P < 0.05) at the same average speed. The results of this experiment were interpreted as indicating that the muscular effort, needed for maintaining balance when walking on a pitching or rolling platform, resulted in a significantly higher work load than similar walking on a stable or a heaving floor, independent of the mode of walking. These results explain in part the increased fatigue observed when a task is performed on a moving platform. Accepted: 3 October 1997  相似文献   

13.
Water permeability of the plasma membrane (PM) and the vacuolar membrane (VM) is important for intracellular and transcellular water movement in plants, because mature plant cells have large central vacuoles. We have developed a new method for measuring the osmotic water permeability of the PM and VM (P f1 and P f2, respectively) in individual plant cells. Here, the theoretical basis and procedure of the method are discussed. Protoplasts isolated from higher plant tissues are used to measure P f1 and P f2. Because of the semi-permeability (selective permeability) of cellular membranes, protoplasts swell or shrink under hypotonic or hypertonic conditions. A theoretical three-compartment model is presented for simulating time-dependent volume changes in the vacuolar and cytoplasmic spaces in a protoplast during osmotic excursions. The model describes the theoretical relationships between P f1, P f2 and the bulk osmotic water permeability of protoplasts (P f(bulk)). The procedure for measuring the osmotic water permeability is: (1) P f(bulk) is calculated from the time when half of the total change in protoplast volume is completed, by assuming that the protoplast has a single barrier to water movement across it (two-compartment model); (2) P f2 of vacuoles isolated from protoplasts is obtained in the same manner; and (3) P f1 is determined from P f(bulk) and P f2 according to the three-compartment model. The theoretical relationship between P fl (m s−1) and L Pl (hydraulic conductivity, l=1, 2) (m s−1 Pa−1) is also discussed. Electronic supplementary material Supplementary material is available in the online version of this article at and is accessible for authorised users. Tsuneo Kuwagata and Mari Murai-Hatano contributed equally to the paper.  相似文献   

14.
Knoche M  Peschel S  Hinz M  Bukovac MJ 《Planta》2000,212(1):127-135
Water conductance of the cuticular membrane (CM) of mature sweet cherry fruit (Prunus avium L. cv. Sam) was investigated by monitoring water loss from segments of the outer pericarp excised from the cheek of the fruit. Segments consisted of epidermis, hypodermis and several cell layers of the mesocarp. Segments were mounted in stainless-steel diffusion cells with the mesocarp surface in contact with water, while the outer cuticular surface was exposed to dry silica (22 ± 1 °C). Conductance was calculated by dividing the amount of water transpired per unit area and time by the difference in water vapour concentration across the segment. Conductance values had a log normal distribution with a median of 1.15 × 10−4 m s−1 (n=357). Transpiration increased linearly with time. Conductance remained constant and was not affected by metabolic inhibitors (1 mM NaN3 or 0.1 mM carbonylcyanide m-chlorophenylhydrazone) or thickness of segments (range 0.8–2.8 mm). Storing fruit (up to 42 d, 1 °C) used as a source of segments had no consistent effect on conductance. Conductance of the CM increased from cheek (1.16 ± 0.10 × 10−4 m s−1) to ventral suture (1.32 ± 0.07 × 10−4 m s−1) and to stylar end (2.53 ± 0.17 × 10−4 m s−1). There was a positive relationship (r2=0.066**; n=108) between conductance and stomatal density. From this relationship the cuticular conductance of a hypothetical astomatous CM was estimated to be 0.97 ± 0.09 × 10−4 m s−1. Removal of epicuticular wax by stripping with cellulose acetate or extracting epicuticular plus cuticular wax by dipping in CHCl3/methanol increased conductance 3.6- and 48.6-fold, respectively. Water fluxes increased with increasing temperature (range 10–39 °C) and energies of activation, calculated for the temperature range from 10 to 30 °C, were 64.8 ± 5.8 and 22.2 ± 5.0 kJ mol−1 for flux and vapour-concentration-based conductance, respectively. Received: 23 March 2000 / Accepted: 28 July 2000  相似文献   

15.
Apoplastic transport across young maize roots: effect of the exodermis   总被引:27,自引:0,他引:27  
The uptake of water and of the fluorescent apoplastic dye PTS (trisodium 3-hydroxy-5,8,10-pyrenetrisulfonate) by root systems of young maize (Zea mays L.) seedlings (age: 11–21 d) has been studied with plants which either developed an exodermis (Casparian band in the hypodermis) or were lacking it. Steady-state techniques were used to measure water uptake across excised roots. Either hydrostatic or osmotic pressure gradients were applied to induce water flows. Roots without an exodermis were obtained from plants grown in hydroponic culture. Roots which developed an exodermis were obtained using an aeroponic (=mist) cultivation method. When the osmotic concentration of the medium was varied, the hydraulic conductivity of the root (Lp r in m3 · m−2 · MPa−1 · s−1) depended on the osmotic pressure gradient applied between root xylem and medium. Increasing the gradient (i.e. decreasing the osmotic concentration of the medium; range: zero to 40 mM of mannitol), increased the osmotic Lp r. In the presence of hydrostatic pressure gradients applied by a pressure chamber, root Lp r was constant over the entire range of pressures (0–0.4 MPa). The presence of an exodermis reduced root Lp r in hydrostatic experiments by a factor of 3.6. When the osmotic pressure of the medium was low (i.e. in the presence of a strong osmotic gradient between xylem sap and medium), the presence of an exodermis caused the same reduction of root Lp r in osmotic experiments as in hydrostatic ones. However, when the osmotic concentration of the medium was increased (i.e. the presence of low gradients of osmotic pressure), no marked effect of growth conditions on osmotic root Lp r was found. Under these conditions, the absolute value of osmotic root Lp r was lower by factors of 22 (hydroponic culture) and 9.7 (aeroponic culture) than in the corresponding experiments at low osmotic concentration. Apoplastic flow of PTS was low. In hydrostatic experiments, xylem exudate contained only 0.3% of the PTS concentration of the bathing medium. In the presence of osmotic pressure gradients, the apoplastic flow of PTS was further reduced by one order of magnitude. In both types of experiments, the development of an exodermis did not affect PTS flow. In osmotic experiments, the effect of the absolute value of the driving force cannot be explained in terms of a simple dilution effect (Fiscus model). The results indicate that the radial apoplastic flows of water and PTS across the root were affected differently by apoplastic barriers (Casparian bands) in the exodermis. It is concluded that, unlike water, the apoplastic flow of PTS is rate-limited at the endodermis rather than at the exodermis. The use of PTS as a tracer for apoplastic water should be abandoned. Received: 9 October 1997 / Accepted: 5 February 1998  相似文献   

16.
The biodegradation of tributyl phosphate (Bu3-P, TBP), releasing phosphate at a high enough concentration locally to precipitate uranium from solution, was demonstrated by a mixed culture consisting primarily of pseudomonads. The effect of various parameters on Bu3-P biodegradation by growing cells is described. Growth at the expense of Bu3-P as the carbon and phosphorus source occurred over a pH range from 6.5 to 8, and optimally at pH 7. Bu3-P biodegradation was optimal at 30 °C, reduced at 20 °C and negligible at 4 °C and 37 °C. Incorporation of Cu or Cd inhibited, and Ni, Co and Mn reduced its degradation. Inorganic phosphate (above 10 mM) and kerosene (up to 1 g/l) reduced Bu3-P biodegradation significantly, but nitrate had no effect. Sulphate (10–100 mM) was inhibitory. When pregrown biomass was used the fastest rates of tributyl and dibutyl phosphate biodegradation were 25 μmol h−1 mg protein−1 and 37 μmol h−1 mg protein−1 respectively. Microcarrier-immobilised biomass decontaminated uranium-bearing acid mine waste water by uranium phosphate precipitation at the expense of Bu3-P hydrolysis in the presence of 35 mM SO4 2−. At pH 4.5, 79% of the UO2 2+ was removed at a flow rate of 1.4 ml/h on a 7-ml test column. Received: 2 June 1997 / Received revision: 15 September 1997 / Accepted: 19 September 1997  相似文献   

17.
A group of 12 healthy non-smoking men [mean age 22.3 (SD 1.1) years], performed an incremental exercise test. The test started at 30 W, followed by increases in power output (P) of 30 W every 3 min, until exhaustion. Blood samples were taken from an antecubital vein for determination of plasma concentration lactate [La]pl and acid-base balance variables. Below the lactate threshold (LT) defined in this study as the highest P above which a sustained increase in [La]pl was observed (at least 0.5 mmol · l−1 within 3 min), the pulmonary oxygen uptake (O2) measured breath-by-breath, showed a linear relationship with P. However, at P above LT [in this study 135 (SD 30) W] there was an additional accumulating increase in O2 above that expected from the increase in P alone. The magnitude of this effect was illustrated by the difference in the final P observed at maximal oxygen uptake (O2max) during the incremental exercise test (P max,obs at O2max) and the expected power output at O2max(P max,exp at O2max) predicted from the linear O2-P relationship derived from the data collected below LT. The P max,obs at O2max amounting to 270 (SD 19) W was 65.1 (SD 35) W (19%) lower (P<0.01) than the P max,exp at O2max . The mean value of O2max reached at P max,obs amounted to 3555 (SD 226) ml · min−1 which was 572 (SD 269) ml · min−1 higher (P<0.01) than the O2 expected at this P, calculated from the linear relationship between O2 and P derived from the data collected below LT. This fall in locomotory efficiency expressed by the additional increase in O2, amounting to 572 (SD 269) ml O2 · min−1, was accompanied by a significant increase in [La]pl amounting to 7.04 (SD 2.2) mmol · l−1, a significant increase in blood hydrogen ion concentration ([H+]b) to 7.4 (SD 3) nmol · l−1 and a significant fall in blood bicarbonate concentration to 5.78 (SD 1.7) mmol · l−1, in relation to the values measured at the P of the LT. We also correlated the individual values of the additional O2 with the increases (Δ) in variables [La]pl and Δ[H+]b. The Δ values for [La]pl and Δ[H+]b were expressed as the differences between values reached at the P max,obs at O2max and the values at LT. No significant correlations between the additional O2 and Δ[La]pl on [H+]b were found. In conclusion, when performing an incremental exercise test, exceeding P corresponding to LT was accompanied by a significant additional increase in O2 above that expected from the linear relationship between O2 and P occurring at lower P. However, the magnitude of the additional increase in O2 did not correlate with the magnitude of the increases in [La]pl and [H+]b reached in the final stages of the incremental test. Accepted: 30 October 1997  相似文献   

18.
Changes in red blood cellsize, deformability, andosmotic fragility are indicators of altered condition and/or altered regulatory processes at the whole cell and membrane levels. An agent, such as HgCl2, that brings about specific changes of this kind can therefore serve as a selective probe of such cell condition and regulatory state. Conversely, for a health-threatening agent “active” in this way, the cell-membrane responses serve to clarify the more fundamental bases of its toxicity, as well as to permit identification and characterization of its early and low-level actions on living systems. Taking advantage of recent advances in the technique of “resistive pulse spectroscopy,” we present a coordinated study of these three interrelated biophysical properties for the interactions of HgCl2 with human red cells. We thereby are able to extend previous studies of this kind into domains of shorter time (instantaneous exposures), lower level exposures (down to 10−9 M, well below the level of acute human toxicity), as well as to additional kinds of responses (e.g., “dynamic osmotic hemolysis”). For conditions ranging from 10−4 to 10−9 M in HgCl2, for instantaneous to 90-min-incubated exposures, for medium osmolarities from 120 to 300, the matrix of observed cell responses includes relative swelling as well as shrinkage, changes in deformability, and both enhancement of and protection against osmotic hemolysis. Some unexpected short-term effects of time and temperature of storage of blood cell stock samples, with respect to increasing and decreasing osmotic fragility, are also reported. These apparently disparate results are interpreted in terms of mercury interactions with cell and membrane SH groups, and a reasonable rationale is presented for most of the responses in terms of disruption of passive and active Na+−K+, gradient controls, plus interactions with cellular proteins.  相似文献   

19.
Evaporation of water from the cell surface of the internode ofChara corallina was not affected by HgCl2 which is known to inhibit water channels. This makes a sharp contrast to the fact that most of osmotically driven water transport is inhibited by HgCl2. Also in radish hypocotyls whose epidermis had been peeled off, evaporation of water was not inhibited by HgCl2, while osmotic water transport was significantly inhibited. The cell wall tube was prepared by squeezing out the content of theChara internode. The rate of evaporation from the cell wall tube filled with 150 mM KCl was almost equal to that from the living cell. The apparent hydraulic conductivity of the cell calculated from evaporation rate was found to be 1–2×10−3 pm s−1 Pa−1 which is about 1/1000 times the hydraulic conductivity of the plasma membrane (Lp) and 1/40 times the Lp under maximal inhibition with HgCl2. It is concluded that under the relative humidity of 53–70% the rate of evaporation of water from the cell surface is limited by the rate of evaporation from the cell wall which is so low that the loss of water can be supplemented without delay from the cell interior across the plasma membrane even when water channels are completely closed.  相似文献   

20.
A laboratory study investigated the metabolic physiology, and response to variable periods of water and sodium supply, of two arid-zone rodents, the house mouse (Mus domesticus) and the Lakeland Downs short-tailed mouse (Leggadina lakedownensis) under controlled conditions. Fractional water fluxes for M. domesticus (24 ± 0.8%) were significantly higher than those of L. lakedownensis (17 ± 0.7%) when provided with food ad libitum. In addition, the amount of water produced by M. domesticus and by L. lakedownensis from metabolic processes (1.3 ± 0.4 ml · day−1 and 1.2 ± 0.4 ml · day−1, respectively) was insufficient to provide them with their minimum water requirement (1.4 ± 0.2 ml · day−1 and 2.0 ± 0.3 ml · day−1, respectively). For both species of rodent, evaporative water loss was lowest at 25 °C, but remained significantly higher in M. domesticus (1.1 ± 0.1 mg H2O · g−0.122 · h−1) than in L. lakedownensis (0.6 ± 0.1 mg H2O · g−0.122 · h−1). When deprived of drinking water, mice of both species initially lost body mass, but regained it within 18 days following an increase in the amount of seed consumed. Both species were capable of drinking water of variable saline concentrations up to 1 mol · l−1, and compensated for the increased sodium in the water by excreting more urine to remove the sodium. Basal metabolic rate was significantly higher in M. domesticus (3.3 ± 0.2 mg O2 · g−0.75 · h−1) than in L. lakedownensis (2.5 ± 0.1 mg O2 · g−0.75 · h−1). The study provides good evidence that water flux differences between M. domesticus and L. lakedownensis in the field are due to a requirement for more water in M. domesticus to meet their physiological and metabolic demands. Sodium fluxes were lower than those observed in free-ranging mice, whose relatively high sodium fluxes may reflect sodium associated with available food. Accepted: 16 August 1999  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