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.
Measurement of the transport parameters that govern the passage of urea and amides across the red cell membrane leads to important questions about transport of water. It had initially been thought that small protein channels, permeable to water and small solutes, traversed the membrane (see Solomon, 1987). Recently, however, very strong evidence has been presented that the 28 kDa protein, CHIP28, found in the red cell membrane, is the locus of the water channel (see Agre et al., 1993). CHIP28 transports water very rapidly but does not transport small nonelectrolytes such as urea. The irreversible thermodynamic parameter, σ i , the reflection coefficient, is a measure of the relationship between the permeability of the solute and that of water. If a solute permeates by dissolution in the membrane, σ i = 1.0; if it permeates by passage through an aqueous channel, σ i < 1.0. For urea, Goldstein and Solomon (1960) found that σurea= 0.62 ± 0.03 which meant that urea crosses the red cell membrane in a water-filled channel. This result and many subsequent observations that showed that σurea < 1.0 are at variance with the observation that CHIP28 is impermeable to urea. In view of this problem, we have made a new series of measurements of σ i for urea and other small solutes by a different method, which obviates many of the criticisms Macey and Karan (1993) have made of our earlier method. The new method (Chen et al., 1988), which relies upon fluorescence of the intracellular dye, fluorescein sulfonate, leads to the corrected value, σurea,corr= 0.64 ± 0.03 for ghosts, in good agreement with earlier data for red cells. Thus, the conclusion on irreversible thermodynamic and other grounds that urea and water share a common channel is in disagreement with the view that CHIP28 provides the sole channel for water entrance into the cell. Received: 6 February 1996/Revised: 20 May 1996  相似文献   

2.
The simultaneous efflux of tritiated water and 14C labelled ethanol from inner epidermal cells of the bulb scale of Allium cepa was measured with a specially designed efflux chamber. It was found that water and ethanol moved essentially independently. Rates of efflux of tritiated water and 14C ethanol were essentially the same in the presence or absence of a simultaneous influx of water. Using the same technique the efflux of tritiated water from the epidermal cells was measured during a simultaneous flow of nonlabelled ethanol. When tritiated water and ethanol moved in opposite directions, the water permeability values became slightly reduced depending upon the concentration of ethanol. When ethanol and tritiated water moved in the same direction, however, no effect on water permeability values could be detected. These results are best explained by the molecular theory of diffusion across lipid bilayer membranes, and are consistent with the above findings of lack of interaction between water and ethanol as they are transported across the cell membrane. In another study, the solute permeability coefficients (Ks) for non-electrolytes such as urea and methyl urea were measured by plasmolyzing the epidermal cells and transferring them to equimolal solutions of urea and methyl urea. This method was also used to measure the reflection coefficient (σ) for these nonelectrolytes. The Ks values for methyl urea were 16 times greater than the ones for urea. The values of σ for both of these solutes, however, were very close to 1. Using the Ks data available in the literature for the subepidermal cells of the Pisum sativum stem basis, the σ values were calculated for malonamide, glycerol, methyl urea, ethyl urea, dimethyl urea, and formamide. Again the Ks values for these nonelectrolytes varied by several orders of magnitude, whereas all σ values were found to be close to 1. These findings point out that σ is an insensitive parameter and that Ks, the solute permeability constant, has to be used for characterizing solute transport through the membrane. The present study shows that fast (e.g. ethanol, formamide) as well as slowly permeating molecules do not interact with water as they are transported across the cell membrane. Aqueous pores for the simultaneous transport of water and solutes, therefore, are absent in the plant cell membranes investigated here.  相似文献   

3.
Knowledge of bovine oocyte plasma membrane permeability characteristics at different developmental stages in the presence of cryoprotective agents (CPAs) is limited. The objective of this study was to determine the oolema hydraulic conductivity (Lp), cryoprotectant permeability (PCPA), and reflection coefficient (σ) for immature (germinal vesicle stage, GV) and in vitro–matured (metaphase II, MII) bovine oocytes. Two commonly used cryoprotective agents, dimethyl sulfoxide (DMSO) and ethylene glycol (EG), were studied. Osmometric studies were performed using a micromanipulator connected to an inverted microscope at 22 ± 2°C. Each oocyte was immobilized via a holding pipette, and osmotically induced volume changes over time (dv/dt) were recorded. The Lp values for GV and MII oocytes in DMSO (LpDMSO) were 0.70 ± 0.06 and 1.14 ± 0.07 μm/min/atm (mean ± SEM) and in EG (LpEG) were 0.50 ± 0.06 and 0.83 ± 0.07 μm/min/atm, respectively. Estimates of PDMSO for GV and MII oocytes were 0.36 ± 0.03 and 0.48 ± 0.03 μm/sec, and PEG values for GV and MII oocytes were 0.22 ± 0.03, 0.37 ± 0.03 μm/sec, respectively. The σ values for GV and MII oocytes in DMSO (σDMSO) were 0.86 ± 0.03 and 0.90 ± 0.04 and in EG (σEG) were 0.94 ± 0.03 and 0.76 ± 0.04, respectively. These data demonstrate that bovine oolema permeability coefficients to water and cryoprotectants change after in vitro maturation. Furthermore, the bovine oocyte PDMSO is higher than the PEG. These results may provide a biophysical basis for developing criteria for choosing optimal CPAs and for minimizing damage during addition and removal of the CPAs. Additionally, these data support the hypothesis that different procedures may be required for optimal cryopreservation of different oocyte developmental stages. Mol. Reprod. Dev. 49:408–415, 1998. © 1998 Wiley-Liss, Inc.  相似文献   

4.
Osmotic responses of maize roots   总被引:16,自引:0,他引:16  
Water and solute relations of excised seminal roots of young maize (Zea mays L) plants, have been measured using the root pressure probe. Upon addition of osmotic solutes to the root medium, biphasic root pressure relaxations were obtained as theoretically expected. The relaxations yielded the hydraulic conductivity Lp r) the permeability coefficient (P sr), and the reflection coefficient (σ sr) of the root. Values of Lp r in these experiments were by nearly an order of magnitude smaller than Lp r values obtained from experiments where hydrostatic pressure gradients were used to induce water flows. The value of P sr was determined for nine different osmotica (electrolytes and nonelectrolytes) which resulted in rather variable values (0.1·10-8–1.7·10-8m·s-1). The reflection coefficient σ sr of the same solutes ranged between 0.3 and 0.6, i.e. σ sr was low even for solutes for which cell membranes exhibit a σ s≈1. Deviations from the theoretically expected biphasic responses occured which may have reflected changes of either P sr or of active pumping induced by the osmotic change. The absolute values of Lp r, P sr, and σ sr have been critically examined for an underestimation by unstirred layer effecs. The data indicate a considerable apoplasmic component for radial movement of water in the presence of hydrostatic gradients and also some solute flow byppassing root protoplasts. In the presence of osmotic gradients, however, there was a substantial cell-to-cell transport of water. Cutting experiments demonstrated that the hydraulic resistance for the longitudinal movement of water was much smaller than for radial transport except for the apical ends of the segments (length=5 to 20 mm). The differences in Lp r as well as the low σ sr values suggest that the simple osmometer model of the root with a single osmotic barrier exhibiting nearly semipermeable properties should be extended for a composite membrane model with hydraulic and osmotic barriers arranged in series and in parallel.  相似文献   

5.
6.
The free energy difference between two states of a molecular system separated by an energy barrier can generally be computed using the technique of umbrella sampling along a chosen reaction coordinate or pathway. The effect of a particular choice of pathway upon the obtained free energy difference is investigated by molecular dynamics simulation of a model system consisting of a glycine dipeptide in aqueous solution. Two different reaction coordinates connecting the so-called C5 and C7 conformations, one involving intramolecular hydrogen bonds and the other involving the peptide ?, ψ angles, are considered. The Gibbs free energy differences ΔG(C5 – C7) are small in both cases, 1.5 ± 1 kJ mol?1 and 2.2 ± 1 kJ mol ?1, respectively. The two different reaction coordinates yield free energy differences that are identical to within their statistical error. It is found that the exchange of solute–solute, solute–water, and water–water hydrogen bonds involves free energy changes of less than kBT, which points at the existence of a multitutde of low free energy pathways connecting the C5 and C7 dipeptide conformations. © 1994 John Wiley & Sons, Inc.  相似文献   

7.
In ruminants, gastrointestinal recycling of urea is acutely enhanced by fibre-rich diets that lead to high ruminal concentration of short chain fatty acids (SCFA), while high ammonia has inhibitory effects. This study attempted to clarify if urea flux to the porcine cecum is similarly regulated. Thirty-two weaned piglets were fed diets containing protein (P) of poor prececal digestibility and fibre (F) at high (H) or low levels (L) in a 2 × 2 factorial design. After slaughter, cecal content was analyzed and the cecal mucosa incubated in Ussing chambers to measure the effect of pH, SCFA and NH4 + on the flux rates of urea, short-circuit current (I sc) and tissue conductance (G t). NH4 + significantly enhanced I sc (from 0.5 ± 0.2 to 1.2 ± 0.1 μEq cm?2 h?1). No acute effects of SCFA or ammonia on urea flux were observed. Tissue conductance was significantly lower in the high dietary fibre groups irrespective of the protein content. Only the HP-LF group emerged as different from all others in terms of urea flux (74 ± 6 versus 53 ± 3 nmol cm?2 h?1), associated with higher cecal ammonia concentration and reduced fecal consistency. The data suggest that as in the rumen, uptake of ammonia by the cecum may involve electrogenic transport of the ionic form (NH4 +). In contrast to findings in the rumen, neither a high fibre diet nor acute addition of SCFA enhanced urea transport across the pig cecum. Instead, a HP-LF diet had stimulatory effects. A potential role for urea recycling in stabilizing luminal pH is discussed.  相似文献   

8.
The resistance of oranges (Citrus sinensis L. Osbeck) and grapefruit (Citrus paradisi Macf.) to ethylene, O2, CO2, and H2O mass transport was investigated anatomically with scanning electron microscope and physiologically by gas exchange measurements at steady state. The resistance of untreated fruit to water vapor is far less than to ethylene, CO2 and O2. Waxing partially or completely plugs stomatal pores and forms an intermittent cracked layer over the surface of fruit, restricting transport of ethylene, O2, and CO2, but not of water; whereas individual sealing of fruit with high density polyethylene films reduces water transport by 90% without substantially inhibiting gas exchange.

Stomata of harvested citrus fruits are essentially closed. However, ethylene, O2 and CO2 still diffuse mainly through the residual stomatal opening where the relative transport resistance (approximately 6,000 seconds per centimeter) depends on the relative diffusivity of each gas in air. Water moves preferentially by a different pathway, probably through a liquid aqueous phase in the cuticle where water conductance is 60-fold greater. Other gases are constrained from using this pathway because their diffusivity in liquid water is 104-fold less than in air.

  相似文献   

9.
10.
Summary Urea and water transport across the toad bladder can be separately activated by low concentrations of vasopressin or 8 Br-cAMP. Employing this method of selective activation, we have determined the reflection coefficient () of urea and other small molecules under circumstances in which the bladder was transporting urea or water. An osmotic method for the determination of was used, in which the ability of a given solute to retard water efflux from the bladder was compared to that of raffinose (=1.0) or water (=0). When urea transport was activated (low concentration of vasopressin), for urea and other solutes was low, (urea,0.08–0.39;acetamide, 0.55; ethylene glycol, 0.60). When water transport was activated (0.1mm 8 Br-cAMP) urea approached 1.0 urea also approached 1.0 at high vasopressin concentrations. In a separate series of studies, urea was determined in the presence of 2×10–5 m KMnO4 in the luminal bathing medium. Under these conditions, when urea transport is selectively blocked, urea rose from a value of 0.12 to 0.89. Thus, permanganate appears to close the urea transport channel. These findings indicate that the luminal membrane channels for water and solutes differ significantly in their dimensions. The solute channels, limited in number, have relatively large radii. They carry a small fraction (approximately 10%) of total water flow. The water transport channels, on the other hand, have small radii, approximately the size of a water molecule, and exclude solutes as small as urea.  相似文献   

11.
The diffusional water permeability (PD) and the pressure filtration coefficient (LP) of isolated larval hindgut cuticle of the fleshfly, Sarcophaga bullata, were measured using tracer techniques coupled with a simple mathematical model system based on equations of non-equilibrium thermodynamics. Data obtained from the model system were matched to experimental tracer data by means of a mathematical optimization scheme. The following parameter values were obtained: PD for tritiated water = 1.02 × 10?6 cm-sec?1, and LP = 9.18 × 10?11 cm3-dyn?1-sec?1. These results are now being used to determine reflection coefficients (σ's) and solute mobilities (ω's) for the cuticle system in an attempt to gain an understanding of the mechanisms controlling solute and water movements across the hindgut wall.  相似文献   

12.
Gastrointestinal parasitism is a global problem for grazing ruminants which can be addressed in a sustainable way through breeding animals to be more resistant to disease. This study estimates the genetic parameters of common and new disease phenotypes associated with natural nematode and coccidian infection in Scottish Blackface sheep to underpin future genetic improvement strategies for parasite control. Data on faecal egg counts (FEC) from different species of strongyle parasites and faecal oocyst counts (FOC) from coccidian parasites were collected on 3-month-old lambs together with a faecal soiling score in the breech area dagginess (DAG) and live weight (LWT). Faecal count data were obtained for Strongyles (FECS), Nematodirus (FECN) and Coccidia (FOC). Data from 3 731 lambs sampled between 2011 and 2017 were included. Faecal egg counts and DAG records were log-transformed prior to analysis. Data were analysed using linear mixed models. Average age at sampling was 92 days with a mean LWT of 24.5 kg. Faecal soiling was not evident in 69% of lambs. Coccidia were the most prevalent parasite (99.5%), while Strongyles and Nematodirus had a prevalence of 95.4% and 72.7%, respectively. Heritability estimates (± SE) were 0.16 ± 0.03, 0.17 ± 0.03, 0.09 ± 0.03, 0.09 ± 0.03 and 0.33 ± 0.04 for FECS, FECN, FOC, DAG and LWT, respectively. Strongyles faecal egg count had a strong and positive genetic correlation with FECN (0.74 ± 0.09) and a moderate positive correlation with FOC (0.39 ± 0.15) while DAG was negatively genetically correlated with LWT (− 0.33 ± 0.15). The significant positive genetic correlations between FECS, FECN and FOC at 3 months of age show that co-selection of sheep for resistance to these different parasites is feasible. Selection for increased resistance to parasite infection is not expected to adversely affect live BW, as no significant antagonistic genetic correlations were found between LWT and FEC. There were significant antagonistic phenotypic and genetic relationships between DAG and LWT being − 0.19 ± 0.02 and − 0.33 ± 0.15, respectively, indicating that the expression of the manifestation of disease in lambs may be a more meaningful indicator of the impact of parasite burden on productivity.  相似文献   

13.
Hydrolysis and absorption of glycylglycine and glycyl-L-leucine as well as absorption of glycine and leucine were studied in chronic experiments on rats with their isolated small intestine loop. Values of the “true” kinetic constants (with taking into account effect of the preepithelial layer) were determined to be as follows: (1) K t = 46.7 ± 4.0 and 2.15 ± 0.59 mM, J max = 0.74 ± 0.15 and 0.16 ± 0.03 μmol min?1 cm?1 (for transport of free glycine and leucine, respectively); (2) K t = 4.4 ± 0.6 and 4.8 ± 0.9 mM, J max = 0.24 ± 0.02 and 0.23 ± 0.02 μmol min?1 cm?1 (for transport of glycylglycine and glycyl-L-leucine, respectively); (3) K M = 5.4 ± 1.0 and 38.2 ± 4.4 mM, V max = 0.09 ± 0.02 and 0.24 ± 0.07 μmol min?1 cm?1 (for membrane hydrolysis of these dipeptides, respectively). According to our calculations, in the wide range of the initial glycylglycine concentrations (2.5–40 mM) a part of the peptide component in its total absorption accounts for 0.77–0.80. In the case of glycyl-L-leucine a part of the peptide component in the total glycine absorption decreases from 0.89 to 0.84, while in the total leucine absorption—from 0.86 to 0.71, the initial dipeptide concentration rising from 5 to 40 mM. The obtained results show that the peptide component prevails in absorption of the studied dipeptides in the rat small intestine, but its role is much lesser than what many authors believe. In the case of glycyl-L-leucine, the peptide component can achieve saturation in the range of high substrate concentrations, its part decreasing essentially to become compared with absorption of free amino acids formed as a result of the dipeptide membrane hydrolysis.  相似文献   

14.
In the present work the coupling under short-circuited conditions between the net Na+-influx across isolated frog skin and the transepithelial transport of water was examined i.e., the short-circuit current (I sc ) and the transepithelial water movement (TEWM) were measured simultaneously. It has been shown repeatedly that the I sc across isolated frog skin is equal to the net transepithelial Na+ transport. Furthermore the coupling between transepithelial uptake of NaCl under open-circuit conditions and TEWM was also measured. The addition of antidiuretic hormone (AVT) to skins incubated under short-circuited conditions resulted in an increase in the I sc and TEWM. Under control conditions I sc was 9.14 ± 2.43 and in the presence of AVT 45.9 ± 7.3 neq cm−2 min−1 (n= 9) and TEWM changed from 12.45 ± 4.46 to 132.8 ± 15.8 nL cm−2 min−1. The addition of the Na+ channel blocking agent amiloride resulted in a reduction both in I sc and TEWM, and a linear correlation between I sc and TEWM was found. The correlation corresponds to that 160 ± 15 (n= 7) molecules of water follow each Na+ across the skin. In another series of experiments it was found that there was a linear correlation between I sc and the increase in apical osmolarity needed to stop the TEWM. The data presented indicate that the observed coupling between the net transepithelial Na+ transport and TEWM is caused by local osmosis. Received: 16 October 1996/Revised: 6 March 1997  相似文献   

15.
The effects of fasting and refeeding on amino acid transport in the perfused rat exocrine pancreas were investigated using a rapid dual tracer dilution technique. Unidirectional amino acid influx (15 s) was quantified (relative to the extracellular tracer d-mannitol) over a wide range of perfusate concentrations in pancreata isolated frm fed and 24 h, 48 h, and 72 h fasted and 72 h fasted and refed (24 h) animals. In fed animals transport of phenylalamine (1–24 mM) and l-serine (1–50 mM) was saturable and weighted non-linear regression analyses of the overall transport indicated an apparent Kt=10±3mM and Vmax=7.0±1.0 μmol/min per g (n = 7) for phenylalanine and Kt=16±3 mM and Vmax=20.6±2.1 μmol/min per g (n = 5) for serine. Fasting animals for 24 h or 48 h did not change the kinetics of either phenylalanine or serine transport. After a 72 h fast the rate of phenylalanine transport (Vmax=15.9±2.9 μmol/min per g, (n = 5) was enhanced whereas the transport affinity (Kt=11±3 mM) remained unaltered. l-Serine transport was essentially unaltered. When 72 h fasted animals were refed for 24 h the Vmax for the phenylalanine transport was reduced to values observed in fed animals. In parallel experiments refeeding had no significant effect on serine transport. Perfusion of pancreata isolated from 72 h fasted animals with bovine insulin (1 mU/ml or 1 μU/ml) did not stimulate either phenylalanine or serine transport. The fasting-induced stimulation of transport may provide a mechanism by which the extracellular supply of essential amino acids as phenylalanine is increased to meet the demands of continued proteolytic and lipolytic enzyme synthesis.  相似文献   

16.
Proteolytic and cataplerotic sources of hepatic glutamine were determined by 2H NMR analysis of urinary phenylacetylglutamine (PAGN) 2H-enrichments in eight healthy subjects after 2H2O and phenylbutyric acid ingestion. Body water enrichment was 0.49±0.03%. PAGN was enriched to lower levels with significant differences between the various glutamine positions. PAGN position 2 enrichment=0.33±0.02%; 3R=0.27±0.02%; 3S=0.27±0.02% and position 4=0.17±0.01%. Position 3R,S enrichments are conditional with the net conversion of citrate to glutamate and are therefore markers of cataplerosis. From the ratio of positions 3R,S to body water enrichment, 55±3% of hepatic glutamine was derived from cataplerosis and 45±3% from proteolysis. In conclusion, enrichment of PAGN 3R,S hydrogens relative to that of body water reflects the contribution of cataplerotic and proteolytic sources to hepatic glutamine.  相似文献   

17.
K Okita  A Teramoto  H Fujita 《Biopolymers》1970,9(6):717-738
A new procedure for evaluating u and σ characterizing σ-helix-forming polypeptides in solution was derived from Nagai's theory for the helix–coil transition of such polymers. Here u is the activity for helix formation from random coil, and σ is the helix initiation parameter. The necessary data are the helical content fN at fixed solvent and temperature as a function of N, where N is the degree of polymerization of the polypeptide sample. Such data were obtained from ORD measurements on a number of fractionated samples of poly-N5-(3-hydroxypropyl)-L -glutamine (PHPG) in mixtures of water and methanol covering the complete range of composition and at various termperatures (5–40°C). When analyzed in terms of the proposed procedure, they yielded values of σ which were in the range (3.2 ± 0.6) × 10?4, substantially independent of solvent composition and temperature. These values were much larger than those obtained recently for σ of poly(β-benzyl-L -aspartate) in m-cresol and in a mixture of chloroform and DCA. The data for [η] and s0 (limiting sedimentation coefficient) as functions of molecular weight indicated that the molecular shape of PHPG in pure methanol is essentially rodlike, whereas that in pure water is not entirely randomly coiled but rather may be regarded as an interrupted helix. These indications were consistent with the results from ORD measurements. When plotted against the corresponding values of fN, the values of [η] and [s0] for PHPG in mixtures of water and methanol of various compositions and temperatures formed smooth composite curves, and we attributed these phenomena to the fact that σ of PHPG was nearly constant under these solvent conditions. Here [s0] stands for a reduced limiting sedimentation coefficient which is equal to the inverse friction factor of the solute molecule.  相似文献   

18.
The influence of cell swelling on cell communication was investigated in cardiomyocytes isolated from the ventricle of adult rats. Measurements of dye coupling were performed in cell pairs using intracellular dialysis of Lucifer Yellow CH. The pipette was attached to one cell of the pair and after a gig ohm seal was achieved, the membrane was ruptured by a brief suction allowing the dye to diffuse from the pipette into the cell. Fluorescence of the dye in the injected as well as in non-dialyzed cell of the pair was continuously monitored. The results indicate that in cell pairs exposed to hypotonic solution the cell volume was increased by about 60% within 35 min and the dye coupling was significantly reduced by cell swelling. Calculation of gap junction permeability (P j) assuming an the intracellular volume accessible to intracellular diffusion of the dye as 12% of total cell volume, showed an average P j value of 0.16 ± 0.04 × 10?4 cm/s (n = 35) in the control and 0.89 ± 1.1 × 10?5 cm (n = 40) for cells exposed to hypotonic solution (P < 0.05). Similar results were found assuming intracellular volumes accessible to the dye of 20 and 30% of total cell volume, respectively. Cell swelling did not change the rate of intracellular diffusion of the dye. The results which indicate that cell volume is an important regulator of gap junction permeability, have important implications to myocardial ischemia and heart failure as well as to heart pharmacology because changes in cell volume caused by drugs and transmitters can impair cell communication with consequent generation of slow conduction and cardiac arrhythmias.  相似文献   

19.
K. Katou  T. Taura  M. Furumoto 《Protoplasma》1987,140(2-3):123-132
Summary The mechanism of water movement across roots is, as yet, not well understood. Some workable black box theories have already been proposed. They, however, assumed unrealistic cell membranes with low values of , or were based on a poor anatomical knowledge of roots. The role of root stele in solute and water transport seems to be especially uncertain. An attempted explanation of the nature of root exudation and root pressure by applying the apoplast canal theory (Katou andFurumoto 1986 a, b) to transport in the root stele is given. The canal equations are solved for boundary conditions based on anatomical and physiological knowledge of the root stele. It is found that the symplast cell membrane, cell wall and net solute transport into the wall apoplast are the essential constituents of the canal system. Numerical analysis shows that the canal system enables the coupled transport of solutes and water into a xylem vessel, and the development of root pressure beyond the level predicted by the osmotic potential difference between the ambient medium and the exudate. Observations on root exudation and root pressure previously reported seem to be explained quite well. It is concluded that the movement of water in the root stele although apparently active is essentially osmotic.Abbreviations J v ex volume exudation per root surface - J0 non-osmotic exudation - Lr overall radial hydraulic conductivity of an excised root - reflection coefficient - Cs difference in the osmotic concentration between the bathing medium and the exudate - R gas constant - T absolute temperature - CK molar concentration of K+ - CCl molar concentration of Cl - Cj molar concentration of ion species j - Pj membrane permeability of ion j - zj valence of ion j - F Faraday constant - Vix intracellular electric potential with reference to the canal  相似文献   

20.
Malate efflux from leaf cells of the Crassulacean acid metabolism plant Kalanchoë daigremontiana Hamet et Perrier was studied using leaf slices submerged in experimental solutions. Leaves were harvested at the end of the dark phase and therefore contained high malate levels. Water potentials of solutions were varied between 0 and −5 bar using mannitol (a slowly permeating solute) and ethylene glycol (a rapidly permeating solute), respectively. Mannitol solutions of water potentials down to −5 bar considerably reduced malate efflux. The slowly permeating solute mannitol reduces both water potential and turgor potential of the cells. The water potential of a mannitol solution of −5 bar is just above plasmolyzing concentration. Malate efflux in ethylene glycol at −5 bar was only slightly smaller than at 0 bar, and much higher than in mannitol at −5 bar. Tissues in rapidly permeating ethylene glycol would have turgor potentials similar to tissues in 0.1 mm CaSO4. The results demonstrate that malate efflux depends on turgor potential rather than on water potential of the cells.  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