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 turgor pressure and water relation parameters were determined in single photoautotrophically grown suspension cells and in individual cells of intact leaves of Chenopodium rubrum using the miniaturized pressure probe. The stationary turgor pressure in suspension-cultured cells was in the range of betwen 3 and 5 bar. From the turgor pressure relaxation process, induced either hydrostatically (by means of the pressure probe) or osmotically, the halftime of water exchange was estimated to be 20±10 s. No polarity was observed for both ex- and endosmotic water flow. The volumetric elastic modulus, , determined from measurements of turgor pressure changes, and the corresponding changes in the fractional cell volume was determined to be in the range of between 20 and 50 bar. increases with increasing turgor pressure as observed for other higher plant and algal cells. The hydraulic conductivity, Lp, is calculated to be about 0,5–2·10–6 cm s–1 bar–1. Similar results were obtained for individual leaf cells of Ch. rubrum. Suspension cells immobilized in a cross-linked matrix of alginate (6 to 8% w/w) revealed the same values for the half-time of water exchange and for the hydraulic conductivity, Lp, provided that the turgor pressure relaxation process was generated hydrostatically by means of the pressure probe. Thus, it can be concluded that the unstirred layer from the immobilized matrix has no effect on the calculation of Lp from the turgor pressure relaxation process, using the water transport equation derived for a single cell surrounded by a large external volume. By analogy, this also holds true for Lp-values derived from turgor pressure changes generated by the pressure probe in a single cell within the leaf tissue. The fair similarity between the Lp-values measured in mesophyll cells in situ and mesophyll-like suspension cells suggests that the water transport relations of a cell within a leaf are not fundamentally different from those measured in a single cell.  相似文献   

2.
The water relations of leaves of Tradescantia virginiana were studied using the miniaturized pressure probe (Hüsken, E. Steudle, Zimmermann, 1978 Plant Physiol. 61, 158–163). Under well-watered conditions cell turgor pressures, P o, ranged from 2 to 8 bar in epidermal cells. In subsidiary cells P o was about 1.5 to 4.5 bar and in mesophyll cells about 2 to 3.5 bar. From the turgor pressure, relaxation induced in individual cells by changing the turgor pressure directly by means of the pressure probe, the half-time of water exchange was measured to be between 3 and 100 s for the epidermal, subsidiary, and mesophyll cells. The volumetric elastic modulus, , of individual cells was determined by changing the cell volume by a defined amount and simultaneously measuring the corresponding change in cell turgor pressure. The values for the elastic modulus for epidermal, subsidiary, and mesophyll cells are in the range of 40 to 240 bar, 30 to 200 bar, and 6 to 14 bar, respectively. Using these values, the hydraulic conductivity, L p, for the epidermal, subsidiary, and mesophyll cells is calculated from the turgor pressure relaxation process (on the basis of the thermodynamics of irreversible processes) to be between 1 and 55·10-7 cm s-1 bar-1. The data for the volumetric elastic modulus of epidermal and subsidiary cells indicate that the corresponding elastic modulus for the guard cells should be considerably lower due to the large volume changes of these cells during opening or closing. Recalculation of experimental data obtained by K. Raschke (1979, Encycl. Plant Physiol. N.S., vol. 7, pp 383–441) on epidermal strips of Vicia faba indicates that the elastic modulus of guard cells of V. faba is in the order of 40–80 bar for closed stomata. However, with increasing stomatal opening, i.e., increasing guard cell volume, decreases. Therefore, in our opinion Raschke's results would indicate a relationship between guard cell volume and which would be inverse to that for plant cells known in the literature. assumes values between 20–40 bar when the guard cell colume is soubled.  相似文献   

3.
Changes in cell turgor pressure have been followed in cells of Microcystis sp. transferred to culture medium containing added NaCl at osmolalities of 30–1,500 mosmol kg-1 ( 74–3,680 kPa). Upon upshock turgor decreased, due to osmotically-induced water loss from the cell. However, partial recovery of turgor was then observed in illuminated cells, with maximum turgor regain in media containing 30–500 mosmol kg-1 NaCl. The lightdependent recovery of turgor pressure was completed within 60 min, with no evidence of further changes in cell turgor up to 24 h. This is the first direct evidence that turgor regulation may occur in a prokaryotic organism. Short-term increases in cell K+ content were also observed upon upshock in NaCl, indicating that turgor regain may involve a turgorsensitive K+ uptake system. Estimation of internal K+ concentration in cells transferred to 250 mosmol kg-1 NaCl showed that changes in cell K+ may account for at least half of the observed turgor regain up to 60 min.  相似文献   

4.
Summary The cell walls of plants and fungi are thought to provide the strength required to resist turgor and thus maintain the integrity and morphology of these cells. However, during growth, walls must undergo rapid expansion which requires them to be plastic and therefore weak. In most tip-growing cells there is an apical concentration of F-actin associated with the rapidly expanding cell wall. Disruption of F-actin in the growing tips of hyphae ofSaprolegnia ferax by a localized irradiation, beginning 2–6 m behind the apex, with actin-selective 270 nm uv light caused the hyphae to burst, suggesting that actin supports the weak apical wall against turgor pressure. Bursting was pH dependent and Ca2+ independent at neutral pH. Hyphae burst in the very tip, where the cell wall is expected to be weakest and actin is most concentrated, as opposed to the lower part of the apical taper where osmotic shock induces bursting when actin is intact. When hyphae were irradiated with a wavelength of light that is less effective at disrupting actin, growth was slowed but they failed to burst, demonstrating that bursting was most likely due to F-actin damage. We conclude that F-actin reinforces the expanding apical wall in growing hyphae and may be the prime stress bearing structure resisting turgor pressure in tip growing cells.Abbreviations RP rhodamine phalloidin - F-actin filamentous actin - EGTA ethylene-glycol-bis-(-amino-ethyl ether) N,N-tetra-acetic acid - PIPES piperazine-N,N-bis-(2-ethanesulfonic acid) - uv ultraviolet  相似文献   

5.
Summary Observations of aperture changes as sucrose is added to the solution bathing epidermal strips ofCommelina communis L. allow calculation of the osmotic changes required to open or close the stomatal pore, for comparison with changes in potassium content. With isolated guard cells, in strips in which all cells other than guard cells have been killed, the internal osmotic changes required are 83 mosmol kg–1 m–1 below 10m aperture, 129 mosmol kg–1 m–1 in the range 10–15 m, and 180 mosmol kg–1 m–1 above 15 m. For opening against subsidiary cell turgor in addition to guard cell turgor, in intact strips with live subsidiary and epidermal cells, these figures should each be increased by about 33 mosmol kg–1 m–1. A change in subsidiary cell turgor is magnified in its effects on the water relations of the guard cell by a factor greater than 3.7 for equal changes in the water potential of the two cells, or greater than 4.7 at constant volume of the guard cell.  相似文献   

6.
A micromanipulation method is described for measuring the bursting forces of bacteria and relating them to cell size. At a compression speed of 6.2 m s–1, bursting forces of three samples of rapidly growing Staphylococcus epidermis from a batch culture varied from 3 to 34 N with an average value of 13.8 N (standard error 0.8 N). Escherichia coli grown in continuous culture at a specific growth rate of 0.5 h–1 had bursting forces varying from 1 to 9 N with an average value of 3.6 N (standard error 0.4 N). In squeeze-hold experiments, force relaxation was observed, which was attributed to water loss from the cells, or viscoelasticity, or both. At high compression speed, such as 6.2 m s–1, this relaxation could be neglected. Micromanipulation strength measurements might be used in studies of cell mechanical disruption and of the dependence of cell strength on cell physiology.  相似文献   

7.
Water relation parameters of leaf cells of the aquatic plant Elodea densa have been measured using the pressure probe. For cells in both the upper and lower epidermis it was found that the elastic modulus () and the hydraulic conductivity (Lp) were dependent on cell turgor (P). Lp was (7.8±5.5)·10-7 cm s-1 bar-1 (mean±SD; n=22 cells) for P>4 bar in cells of the upper epidermis and was increasing by a factor of up to three for P0 bar. No polarity of water movement or concentration dependence of Lp was observed. For cells of the lower epidermis the Lp-values were similar and the hydraulic conductivity also showed a similar dependence on turgor. No wall ingrowth or wall labyrinths (as in transfer cells) could be found in the cells of the lower epidermis. The elastic modulus () of cells of the upper epidermis could be measured over the whole pressure range (P=0–7 bar) by changing the osmotic pressure of the medium. increased linearly with increasing turgor and ranged between 10 and 150 bar. For cells of the lower epidermis the dependence of on P was similar, although the pressure dependence could not be measured on single cells. The Lp-values are compared with literature data obtained for Elodea by a nuclear magnetic resonance (NMR)-technique. The dependence of Lp on P is discussed in terms of pressure dependent structural changes of the cell membranes and interactions between solute and water transport.Abbreviations P cell turgor pressure - Lp hydraulic conductivity - volumetric elastic modulus - T 1/2 half-time of water exchange of individual cell  相似文献   

8.
The mechanical properties of a sample of baker's yeast cells were measured by micromanipulation. The relationship between the force required to burst a single cell and its corresponding diameter was established. For stationary phase cells, the compressive force required to burst a cell varied between 55 and 175N, with a mean value of 101 ± 2N. This is a substantial force compared to that required to burst a single mammalian cell (1.5–4.5N), which presumably reflects the lack of a cell wall of the latter. From measurements on 120 cells, there was no significant dependence of bursting force on yeast cell size. The micromanipulation method will be valuable for studying the dependence of mechanical properties of yeast cells on fermentation conditions, and the consequential effects of their behaviour in process disruption operations. © Rapid Science Ltd. 1998  相似文献   

9.
Summary The marine algaValonia macrophysa an inhabitant of shallow subtropical waters, is subjected to sudden dilutions of external seawater during rain showers. This study describes the mechanisms involved in turgor pressure regulation following acute hyposmotic shock. Turgor regulation is 88% effective and complete within 4 hr following hyposmotic shocks of up to –10 bar. Loss of vacuolar K+, Na+ and Cl accounts for the decrease in vacuolar osmotic pressure associated with turgor regulation. A novel mechanism of turgor regulation is exhibited byValonia macrophysa given hyposmotic shocks greater than about –4 bar. Such an osmotic shock causes cell wall tension to increase above a critical value of about 6×105 dyne/cm, whereupon the protoplasm ruptures and the cell wall stretches irreversibly at a localized site. The protoplasm rupture is suggested by (1) a large abrupt increase in K+ efflux (as measured by86Rb+), (2) a rapid decrease in turgor pressure as measured with a pressure probe, and (3) sudden depolarization of the vacuole potential. Evidence for an increase in cell wall permeability includes efflux from the vacuole of dextran (mol wt 70,000), which normally has a very low cell wall permeability, and scanning electron micrographs which show a trabeculated scar area in the cell wall. This mechanism of turgor regulation is physiologically important because 98% of the cells regained normal growth rate and turgor following acute osmotic shock.  相似文献   

10.
The yielding properties of the cell wall, irreversible wall extensibility (m) and yield threshold (Y), are determined for stage I sporangiophores of Phycomyces blakesleeanus from in-vivo creep experiments, and compared to the values of m and Y previously determined for stage IVb sporangiophores using the same pressureprobe method (Ortega et al., 1989, Biophys. J. 56, 465). In either stage the sporangiophore enlarges (grows) predominately in length, in a specific region termed the growing zone, but the growth rates of stage I (5–20 urn · min–1) are smaller than those of stage IVb (30–70 m · min–1). The results demonstrate that this difference in growth rate is the consequence of a smaller magnitude of m for stage I sporangiophores; the obtained values of P (turgor pressure), Y, and P-Y (effective turgor for irreversible wall extension) for stage I sporangiophores are slightly larger than those of stage IVb sporangiophores. Also, it is shown that the magnitude of m for the stage I sporangiophore is regulated by altering the length of the growing zone, Lg. A relationship between m and Lg is obtained which can account for the difference between values of m determined for stage I and stage IVb sporangiophores. Finally, it is shown that similar changes in the magnitude of m and (which have been used interchangeably in the literature as a measure of irreversible wall extensibility) may not always represent the same changes in the cell-wall properties.Abbreviations and Symbols L length - Lg length of growing zone - m irreversible wall extensibility - P turgor pressure - Y yield threshold - (P-Y) effective turgor for irreversible wall extension - relative irreversible wall extensibility - g relative irreversible wall extensibility of the growing zone (m/Lg) This work was supported by National Science Foundation grant DCB-8801717 to J.K.E. Ortega.  相似文献   

11.
The kinetics of the release of chitinolytic activity (endochitinase EC 3.2.1.14, \-N-acetyglucosaminidase EC 3.2.1.30) by a yeast cell wall lytic Arthrobacter species was studied. The organism was cultivated on yeast cell wall, mycelium of Trichoderma reesei, colloidal chitin, N-acetylglucosamine, glucosamine and mixtures with acetate. With the exception of yeast cell wall, these substrates were used as the sole source of carbon and nitrogen. The growth on colloidal chitin (0.5%) proceeded at a maximum specific growth rate (umax) of 0.23 h–1 and yielded 2700 mU1–1 chitinase. Yeast cell wall and mycelium of T. reesei supported more rapid growth (max = 0.30 h–1 and 0.25 h–1 respectively) but yielded reduced chitinase activity (565 mUl–1 and 700 mUl–1). The growth rate on glucosamine (max = 0.24 h–1) was reduced when this was mixed with acetate (max = 0.12 h–1), whereas the enzyme yield was increased from 720 mUl–1 to 960 mUl–1. The same effect on growth rate was observed with glucose and equimolar mixtures of glucose and acetate, indicating a strong impact of the organic acid on carbohydrate transport or metabolism. The growth of adapted cells on N-acetylglucosamine was comparable to that observed on an equimolar mixture of glucosamine and acetate, indicating that N-acetylglucosamine is rapidly hydrolysed by adapted cells.  相似文献   

12.
Theory predicts that, for growing plant cells isolated from a supply of water, stress relaxation of the cell wall should decrease cell turgor pressure (P) until the yield threshold for cell expansion is reached. This prediction was tested by direct P measurements of pea (Pisum sativum L.) stem cortical cells before and after excision of the growing region and isolation of the growing tissue from an external water supply. Cell P was measured with the micro-pressure probe under conditions which eliminated transpiration. Psychrometric measurements of water potential confirmed the pressureprobe measurements. Following excision, P of the growing cells decreased in 1 h by an average of 1.8 bar to a mean plateau value of 2.8 bar, and remained constant thereafter. Treatment with 10-5 M indole-3-acetic acid or 10-5 M fusicoccin (known growth stimulants) accelerated the rate of P relaxation, whereas various treatments which inhibit growth slowed down or completely stopped P relaxation in apical segments. In contrast, P of basal (nongrowing) segments gradually increased because of absorption of solutes from the cell-wall free space of the tissue. Such solute absorption also occurred in apical segments, but wall relaxation held P at the yield threshold in those segments which were isolated from an external water supply. These results provide a new and rapid method for measuring the yield threshold and they show that P in intact growing pea stems exceeds the yield threshold by about 2 bar. Wall relaxation is shown here to affect the water potential and turgor pressure of excised growing segments. In addition, solute release and absorption upon excision may influence the water potential and turgor pressure of nongrowing excised plant tissues.Abbreviations and symbols IAA indole-3-acetic acid - P turgor pressure - SE standard error of the mean - water potential  相似文献   

13.
M. Hohl  P. Schopfer 《Planta》1992,188(3):340-344
Plant organs such as maize (Zea mays L.) coleoptiles are characterized by longitudinal tissue tension, i.e. bulk turgor pressure produces unequal amounts of cell-wall tension in the epidermis (essentially the outer epidermal wall) and in the inner tissues. The fractional amount of turgor borne by the epidermal wall of turgid maize coleoptile segments was indirectly estimated by determining the water potential * of an external medium which is needed to replace quantitatively the compressive force of the epidermal wall on the inner tissues. The fractional amount of turgor borne by the walls of the inner tissues was estimated from the difference between -* and the osmotic pressure of the cell sap (i) which was assumed to represent the turgor of the fully turgid tissue. In segments incubated in water for 1 h, -* was 6.1–6.5 bar at a i of 6.7 bar. Both -* and i decreased during auxin-induced growth because of water uptake, but did not deviate significantly from each other. It is concluded that the turgor fraction utilized for the elastic extension of the inner tissue walls is less than 1 bar, i.e. less than 15% of bulk turgor, and that more than 85% of bulk turgor is utilized for counteracting the high compressive force of the outer epidermal wall which, in this way, is enabled to mechanically control elongation growth of the organ. This situation is maintained during auxin-induced growth.Abbreviations and Symbols i osmotic pressure of the tissue - 0 external water potential - * water potential at which segment length does not change - IAA indole-3-acetic acid - ITW longitudinal inner tissue walls - OEW outer epidermal wall - P turgor Supported by Deutsche Forschungsgemeinschaft (SFB 206).  相似文献   

14.
K. H. Büchner  U. Zimmermann 《Planta》1982,154(4):318-325
Cells of Halicystis parvula, Acetabularia mediterranea, and Valonia utricularis were immobilized in a cross-linked alginate matrix (4–6% w/w) in order to simulate water-relation experiments in individual cells of higher plant tissues. The immobilization of these cells did not lead to an increase in the mechanical stability of the cell walls. This was demonstrated by measuring the volumetric elastic modulus of the cell wall and its dependence on turgor pressure with the aid of the non-miniaturized pressure probe. In immobilized cells, no changes in the absolute value of the elastic modulus of the cell wall could be detected for any given pressure. At the maximum turgor pressure at which non-immobilized cells normally burst (about 3–7 bar for V. utricularis; depending on cell size, 3 bar for A. mediterranea and 0.9 bar for H. parvula) reversible decreases in the pressure are observed which are succeeded by corresponding pressure increases. This obvervation indicates that coating the cells with the cross-linked matrix protects them from rapid water and turgor pressure loss. Turgor pressure relaxation processes in immobilized cells, which could be induced hydrostatically by means of the pressure probe, yielded accurate values for the half-times of water exchange and for the hydraulic conductivity of the cell membrane. The results demonstrate that the water transport equations derived for single cells in a large surrouding medium are valid for immobilized cells, so that any influence exerted by the unstirred layer which is caused by the presence of the cross-linked matrix can be ignored in the calculations. On the other hand, the evaluation of the half-times of water exchange and the hydraulic conductivity from turgor pressure relaxation processes, which have been induced osmotically, only yields correct values under certain circumstances. The model experiments presented here show, therefore, that the correct Lp-value for an individual cell in a higher plant tissue can probably only be obtained presently by using the pressure probe technique rather than the osmotic method. The results are also discussed in relation to the possible applications of immobilized cells and particularly of immobilized micro-organisms in catalytic reaction runs on an industrial scale.  相似文献   

15.
The long term shear effects on a hybridoma cell line were studied by the simulation of a hollow fiber perfusion system. Various mechanical/environmental stress conditions were applied and steady state concentrations of live, dead and lysed cells were measured or calculated in a continuous culture. From mathematical modeling, it is shown that inclusion of a lysed cell index (LCI) renders a better fit to the material balance equation at steady state. The specific cell death rate increased with increasing shear force as expected only when the LCI was included. Without the inclusion of the LCI, the calculated specific cell growth rates are about 25–60% of the value when included. The results reported may lend some insight to design improvements since most perfusion devices add shear stresses to the cells in the reactor.List of Symbols b ml/hr continuous culture flow rate - D hr–1 dilution rate (b/V) - m g glucose/109 cells/hr specific maintenance coefficient - S 0 g/l feed substrate concentration - S g/l reactor substrate concentration - t hr time - V ml reactor volume - X + cells/ml live cell concentration - X cells/ml dead cell concentration - X 0 cells/ml lysed cell concentration - Y x/s 109 cells/g glucose cell/substrate yield coefficient - hr–1 specific growth rate - hr–1 specific death rate - hr–1 specific lysis rate - hr–1 specific lysis rate for simultaneous death and lysis  相似文献   

16.
Gas vesicle formation and buoyancy regulation in Pelodictyon phaeoclathratiforme strain BU1 (Green sulfur bacteria) was investigated under various laboratory conditions. Cells formed gas vesicles exclusively at light intensities below 5 mol · m-2 · s-1 in the stationary phase. No effect of incubation temperature or nutrient limitation was observed. Gas space of gas vesicles occupied always less than 1.2% of the total cell volume. A maximum cell turgor pressure of 330 kPa was determined which is comparable to values determined for cyanobacterial species. Since a pressure of at least 485 kPa was required to collapse the weakest gas vesicles in Pelodictyon phaeoclathratiforme, short-term regulation of cell density by the turgor pressure mechanism can be excluded.Instead, regulation of the cell density is accomplished by the cease of gas vacuole production and accumulation of carbohydrate at high light intensity. The carbohydrate content of exponentially growing cells increased with light intensity, reaching a maximum of 35% of dry cell mass above 10 mol · m-2 · s-1. Density of the cells increased concomitantly. At maximum density, protein and carbohydrate together accounted for 62% of the total cell ballast. Cells harvested in the stationary phase had a significantly lower carbohydrate content (8–12% of the dry cell mass) and cell density (1010–1014 kg · m-3 with gas vesicles collapsed) which in this case was independent of light intensity. Due to the presence of gas vesicles in these cultures, the density of cells reached a minimum value of 998.5 kg · m-3 at 0.5 mol · m-2 · s-1.The cell volume during the stationary phase was three times higher than during exponential growth, leading to considerable changes in the buoyancy of Pelodictyon phaeoclathratiforme. Microscopic observations indicate that extracellular slime layers may contribute to these variations of cell volume.  相似文献   

17.
The activity of a -(1-4)-xylan synthetase, a membrane-bound enzymic system, was measured in particulate enzymic preparations (1,000 g and 1,000–100,000 g pellets) obtained from homogenates of cambial cells, differentiating xylem cells and differentiated xylem cells isolated from actively growing trees of sycamore (Acer pseudoplatamus) and poplar (Populus robusta). The specific activity (nmol of xylan formed min–1 mg–1 of protein) as well as the activity calculated on a per cell basis (nmol of xylan formed min–1 cell–1) of this enzymic system, markedly increased as cells differentiate from the vascular cambium to xylem. This increase is closely correlated with the enhanced deposition of xylan occurring during the formation of secondary thickening. The possible control of xylan synthesis during the biogenesis of plant cell wall is discussed.  相似文献   

18.
1. In Lymnaea stagnalis L. (Pulmonata, Basommatophora) the neurons in the osphradium were visualized by staining through the inner right parietal nerve by 5,6-carboxyfluorescein (5,6-CF). Three types of neurons were identified: three large ganglionic cells (GC1-3; 80–100 m), the small putative sensory neurons (SC; 20 m) and very small sensory cells (3–5 m).2. The ganglionic and putative sensory neurons were investigated by whole cell patch-clamp method in current-clamp condition. The three giant ganglionic neurons (GC1-3) located closely to the root of osphradial nerve, had a membrane potential (MP) between –30 and –70 mV and showed tonic or bursting activities. The small putative sensory cells (SCs) scattered throughout the osphradial ganglion, possessed a MP between –25 and –55 mV and showed an irregular firing pattern with membrane oscillations. At resting MP the GC1-3 cells were depolarized and increased the frequency of their firing, while the SCs were hyperpolarized and inhibited by NaCl (10–2 M) and L-aspartate (10–5 M) applied to the osphradium.3. 5-Hydroxytryptamine (5HT, 10–6 M), -aminobutyric acid (GABA; 10–6 M) and the GABAB agonist baclofen (10–6 M) depolarized the neurons GC1-3 and increased their firing frequency. In contrast, on the GC1-3 neurons, acetylcholine (Ach; 10–6 M) and FMRFamide (10–6 M) caused hyperpolarization and cessation of the firing activity. The 5HT effect was blocked by mianserin (10–6 M) but picrotoxin (10–5 M) failed to block the GABA-induced effect on the GC1-3 cells.4. The small putative sensory neurons (SCs) were excited by Ach (10–6 M) and 5HT (10–6 M) but were inhibited by GABA (10–6 M). FMRFamide (10–6 M) had a biphasic response. The Ach effect was blocked by hexamethonium (10–6 M) and tetraethylammonium (10–6 M), indicating the involvement of nicotinic cholinergic receptors.5. The distinct responses of the two populations of osphradial neurons to chemical stimuli and neurotransmitters suggest that they can differently perceive signals from environment and hemolymph.  相似文献   

19.
R. J. Reid  F. A. Smith 《Planta》1992,186(4):558-566
This paper deals with the effect of calcium binding in the cell wall on the measured 45Ca influx in Chara corallina Klein ex Will. esk. R.D. Wood. Calcium in the cell wall was in the range 687–1197 (mol · m–2 compared to the sap which contained only 144–256 mol · m–2. In dilute culture solutions the calcium content of the cell wall was relatively independent of external calcium at concentrations above about 0.1 mol · m–3. The half-times for exchange of calcium from 45Ca-labelled cell walls varied from 45 min at 0.05 mol · m–3 to less than 2 min at 2 mol · m–3. The effectiveness of other cations in displacing calcium from cell walls was in the order La > Zn > Co > Ni > Mg. Rinsing of 45Ca-labelled cell walls in 2 mol · m–3 LaCl3 for 20 min removed more than 99% of the bound 45Ca. However, the residual 45Ca activity in isolated cell walls following La3+ rinsing was similar to that in whole cells. It is concluded that in whole cells 45Ca influx cannot normally be distinguished from extracellular binding of calcium. Methods are described for the measurement of 45Ca fluxes in charophyte cells by isolation of intracellular 45Ca after the uptake period using techniques which avoid contamination from the large amount of tracer bound in the cell wall. At an external calcium concentration of 1 mol · m–3, the plasmalemma influx was approx. 0.2 nmol · m–2 · s–1 of which about half entered the vacuole and half was effluxed back into the external solution. The cytoplasm filled with calcium with a half-time of 40–50 min with an apparent pool size of 50 mmol · m–3. After 2 h the net flux to the cell was almost the same as the vacuolar flux. The fluxes reported are an order of magnitude lower than previously reported calcium fluxes in plants.Abbreviations APW artificial pond water This work was supported by the Australian Research Council. The authors wish to thank Patrick Kee for his skilful technical assistance and Professor E.A.C. MacRobbie, University of Cambridge, UK, and Dr. M. Tester for helpful discussions.  相似文献   

20.
M. Tazawa  K. Kiyosawa 《Protoplasma》1973,78(4):349-364
Summary The mechanism of transcellular osmosis was analyzed on the assumption that the driving force, which is equal to the osmotic pressure of the mannitol solution given to the exosmosis side, is divided into two parts; one causing the inward water flow on the water side, the other causing the outward water flow on the solution side, when each force drives an equal amount of water. Based on this analysis a new procedure was developed to measure the endosmotic and exosmotic water permeabilities of the membranes independently. It involved measurement of volume of water transported transcellularly, change in turgor pressure, and water permeability of the cell wall alone.Experiments following the new procedure revealed that in aNitella internode positioned across a partition wall with equal length both the endosmotic and exosmotic water permeabilities remained constant during transcellular osmosis induced with 0.4M mannitol, at least for the first minute. It was found that the permeability coefficient for endosmosis (3.9 × 10–5 cm sec–1 atm–1) was very much higher than that for exosmosis (1.4× 10–5 cm sec–1 atm–1). Treatment of the endosmotic cell part with 5% ethanol conspicuously decreased the water permeability of the cell on this side down to 1/2.4 the value obtained without ethanol but never affected the permeability on the other side (exosmosis side).This work was supported partly by a Research Grant from the Ministry of Education of Japan.  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