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 15 毫秒
1.
Choi  Woo-Jung  Jin  Seong-Ahi  Lee  Sang-Mo  Ro  Hee-Myong  Yoo  Sun-Ho 《Plant and Soil》2001,235(1):1-9
The tolerance of 24 genotypes of barley was assessed by estimating their survival in saline conditions either in a glasshouse or in a controlled environment cabinet. Two cultivars, sensitive Triumph and resistant Gerbel, were picked for further study, which involved sequential harvesting of plants grown in a range of salinities. After about one month in 200 mol m–3 sodium chloride, the sodium concentration in the roots and shoots of the sensitive Triumph was about 1.5 times that in the roots of resistant Gerbel. The addition of Na to the root medium reduced the potassium transport to the shoot in Triumph to a much greater extent than in Gerbel, so the K:Na ratio of Gerbel was twice that for Triumph, when averaged over all treatments and harvests. The sodium, potassium and chloride concentrations within the major subcellular compartments of the cortical cells of roots of Triumph and Gerbel were determined by X-ray microanalysis following freeze-substitution and dry-sectioning. The mean cytoplasmic sodium concentration (245 mol m–3 analysed volume) in Triumph grown in 200 mol m–3 NaCl for 15 d was almost 1.4 times greater than that in the resistant Gerbel: the potassium concentration in Gerbel showed a lower reduction than did that of Triumph. Another major difference between the two cultivars was the higher concentrations of sodium and chloride in the cell walls of Triumph than Gerbel: the sodium concentration in the cortical cell walls of the salt-sensitive cultivar was about 1.75 times that in the more salt-resistant cultivar. The exchange capacity of the cell walls of Gerbel was greater than that of Triumph. We hypothesise that ion transport to the shoot reflects cytosolic ion concentrations, with a more sensitive cultivar having a higher sodium concentration in its cytoplasm than a more resistant variety. It is noteworthy that the difference in the K:Na ratio between the shoots of Gerbel and Triumph after 15 days of exposure to 200 mol m–3 NaCl was similar to the difference in their symplastic K:Na ratios.  相似文献   

2.
D. M. R. Harvey 《Planta》1985,165(2):242-248
Zea mays is a salt-sensitive crop species which in saline (100 mol m-3 NaCl) conditions suffers considerable growth reduction correlated with elevated Na+ and Cl- concentration within the leaves. To increase understanding of the regulation of ion uptake and transport by the roots in saline conditions, ion concentrations within individual root cortical cells were determined by X-ray microanalysis. There was variation in Na+, K+ and Cl- distributions among individual cells, which could not be correlated with their spatial position in the roots. Generally, however, in response to saline growth conditions (100 mol m3 NaCl) Na+ and Cl- were mostly localized in the vacuoles, although their concentrations were also sometimes increased in the cytoplasm and cell walls. The concentration of K+ in the cytoplasm was usually maintained at a level (mean 79 mol m-3) compatible with the biochemical functions ascribed to this ion.Abbreviation (T)AEM (Transmission) analytical electron microscopy  相似文献   

3.
The changes in turgor pressure that accompany the mobilisation of sucrose and accumulation of salts by excised disks of storage-root tissue of red beet (Beta vulgaris L.) have been investigated. Disks were washed in solutions containing mannitol until all of their sucrose had disappeared and then were transferred to solutions containing 5 mol·m-3 KCl+5 mol·m-3 NaCl in addition to the mannitol. Changes in solute contents, osmotic pressure and turgor pressure (measured with a pressure probe) were followed. As sucrose disappeared from the tissue, reducing sugars were accumulated. For disks in 200 mol·m-3 mannitol, the final reducing-sugar concentration equalled the initial sucrose concentration so there was no change in osmotic pressure or turgor pressure. At lower mannitol concentrations, there was a decrease in tissue osmotic pressure which was caused by a turgor-driven leakage of solutes. At concentrations of mannitol greater than 200 mol·m-3, osmotic pressure and turgor pressure increased because reducing-sugar accumulation exceeded the initial sucrose concentration. When salts were provided they were absorbed by the tissue and reducing-sugar concentrations fell. This indicated that salts were replacing sugars in the vacuole and releasing them for metabolism. The changes in salf and sugar concentrations were not equal because there was an increase in osmotic pressure and turgor pressure. The amount of salt absorbed was not affected by the external mannitol concentration, indicating that turgor pressure did not affect this process. The implications of the results for the control of turgor pressure during the mobilisation of vacuolar sucrose are discussed.To whom correspondence should be addressed.  相似文献   

4.
Uptake of 3H-labelled (±)-abscisic acid (ABA) into isolated barley (Hordeum vulgare L.) epidermal cell protoplasts (ECP) was followed over a range of pH values and ABA concentrations. The present results show that ABA uptake is not always linearly correlated with the external concentration of undissociated ABA (ABAH). At pH 7.25, ABA uptake exhibited saturation kinetics with an apparent K m value of 75 mmol·m–3 to tal ABA. This saturable transport component was inhibited by pretreating the protoplasts with 1 mol·m–3 p-chloromercuribenzenesulfonic acid at pH 8.0, conditions that minimized the uptake of this acid sulfhydryl reagent. Moreover, the rate of (±)-[3]HABA uptake was reduced by addition of 0.1 mol·m–3 (±)-ABA to 41%, whereas the same concentration of (±)-ABA was approximately half as effective (46% of the inhibitory effect). Thus, it was concluded that only (±)-ABA competes for an ABA carrier that is located in the epidermal cell plasma membrane. The permeability of the epidermal cell plasma membrane was studied by performing a Collander analysis. At pH 6 the overall plasma-membrane permeability of epidermal cells was similar to that of guard cells but was about two times higher than that of mesophyll cells.Abbreviations ABA abscisic acid - ABA anion of ABA - ABAH undissociated ABA - 2,4-D 2,4-dichlorophenoxyacetic acid - DMO 5,5-dimethyloxazolidine-2,4-dione - ECP deepidermal cell protoplast - Kr partition coefficient - Mr relative molecular mass - NEM N-ethylmaleimide - PCMBS p-chloromercuriben zenesulfonic acid - Ps permeability coefficient We are grateful to Barbara Dierich for expert technical assistance, to Prof. H. Gimmler (Lehrstuhl für Botanik I, Universität Würzburg, FRG) for helpful discussions and to the Deutsche Forschungsgemeinschaft (SFB 251, TP 3) for financial support.  相似文献   

5.
Abscisic acid (ABA) was shown to influence turgor pressure and growth in wheat (Triticum aestivum L.) roots. At a concentrations of 25 mmol·m-3, ABA increased the turgor pressure of cells located within 1 cm of the tip by up to 450 kPa. At 4 to 5 cm from the root tip this concentration of ABA reduced the turgor pressure of peripheral cells (epidermis and the first few cortical cell layers) to zero or close to zero while that of the inner cells was increased. Increases in sap osmolality were dependent on the concentration of ABA and the effect saturated at 5 mmol·m-3 ABA. The increase in osmolality took about 4 h and was partly the result of reducing-sugar accumulation. Levels of inorganic cations were not affected by ABA. Root growth was inhibited at ABA concentrations that caused a turgor-pressure increase. The results show that while ABA can affect root cell turgor pressures, this effect does not result in increased root growth.Abbreviation ABA abscisic acid  相似文献   

6.
Nitrate-selective microelectrodes were used to measure intracellular nitrate concentrations (as activities) in epidermal and cortical cells of roots of 5-d-old barley (Hordeum vulgare L.) seedlings grown in nutrient solution containing 10 mol · m–3 nitrate. Measurements in each cell type grouped into two populations with mean (±SE) values of 5.4 ± 0.5 mol · m–3 (n=19) and 41.8 ± 2.6 mol · m–3 (n = 35) in epidermal cells, and 3.2 ± 1.2 mol · m–3 (n = 4) and 72.8 ± 8.4 mol · m–3 (n = 13) in cortical cells. These could represent the cytoplasmic and vacuolar nitrate concentrations, respectively, in each cell type. To test this hypothesis, a single-cell sampling procedure was used to withdraw a vacuolar sap sample from individual epidermal and cortical cells. Measurement of the nitrate concentration in these samples by a fluorometric nitrate-reductase assay confirmed a mean vacuolar nitrate concentration of 52.6 ± 5.3 mol · m–3 (n = 10) in epidermal cells and 101.2 ± 4.8 mol · m–3 (n = 44) in cortical cells. The nitrate-reductase assay gave only a single population of measurements in each cell type, supporting the hypothesis that the higher of the two populations of electrode measurements in each cell type are vacuolar in origin. Differences in the absolute values obtained by these methods are probably related to the fact that the nitrate electrodes were calibrated against nitrate activity but the enzymic assay against concentration. Furthermore, a 28-h time course for the accumulation of nitrate measured with electrodes in epidermal cells showed the apparent cytoplasmic measurements remained constant at 5.0 ± 0.7 mol · m–3, while the vacuole accumulated nitrate to 30–50 mol · m–3. The implications of the data for mechanisms of nitrate transport at the plasma membrane and tonoplast are discussed.Symbol n 2 Chi-squared with n degrees of freedom R.-G.Z. was awarded a Sino-British Friendship Scholarship sponsored by the British Council and H.-W.K. was supported by an AFRC Linked Research Grant to A.D.T for collaboration with R.A.L. We wish to thank Dr. K. Goulding for advice on ion chromatography, Dr. K. Moore for assistance with statistical analysis and Dr. J.H. Williams for advice on the microsample analysis.  相似文献   

7.
Photoinhibition of photosynthesis was induced in intact leaves of Phaseolus vulgaris L. grown at a photon flux density (PFD; photon fluence rate) of 300 mol·m-2·s-1, by exposure to a PFD of 1400 mol·m-2·s-1. Subsequent recovery from photoinhibition was followed at temperatures ranging from 5 to 35°C and at a PFD of either 20 or 140 mol·m-2·s-1 or in complete darkness. Photoinhibition and recovery were monitored mainly by chlorophyll fluorescence emission at 77K but also by photosynthetic O2 evolution. The effects of the protein-synthesis inhibitors, cycloheximide and chloramphenicol, on photoinhibition and recovery were also determined. The results demonstrate that recovery was temperature-dependent with rates slow below 15°C and optimal at 30°C. Light was required for maximum recovery but the process was light-saturated at a PFD of 20 mol·m-2·s-1. Chloramphenicol, but not cycloheximide, inactivated the repair process, indicating that recovery involved the synthesis of one or more chloroplast-encoded proteins. With chloramphenicol, it was shown that photoinhibition and recovery occurred concomitantly. The temperature-dependency of the photoinhibition process was, therefore, in part determined by the effect of temperature on the recovery process. Consequently, photoinhibition is the net difference between the rate of damage and the rate of repair. The susceptibility of chilling-sensitive plant species to photoinhibition at low temperatures is proposed to result from the low rates of recovery in this temperature range.Abbreviations and symbols Da Dalton - Fo, Fm, Fv instantaneous, maximum, variable fluorescence emission - PFD photon flux density - PSII photosystem II - photon yield C.I.W.-D.P.B. Publication No. 871  相似文献   

8.
In mesophyll cells of species with a symplasmic (Ocimum basilicum, Catharanthus roseus, Magnolia denudata) or an apoplasmic (Vicia faba, Impatiens walleriana, Bellis perennis) minor-vein configuration, membrane depolarizations in response to 20 or 200 mol·m–3 raffinose and sucrose were measured. Ageing period and resting potential marginally affected the degree of depolarization. The symplasmic species showed similar depolarization responses to 20 and 200 mol·m–3 sucrose or raffinose. In the apoplasmic species, depolarization increased statistically significantly from 20 to 200 mol·m–3 sucrose, whereas the depolarization response to raffinose was equal at both concentrations. In the apoplasmic species, moreover, the depolarization response to raffinose was significantly weaker than to sucrose at all concentrations. A major difference between symplasmic and apoplasmic species seems to lie in the scantiness of raffinose carriers in the mesophyll plasma membrane of species with the apoplasmic mode of phloem loading.Abbreviations 20R(200R) 20(200) mol·m–3 raffinose - 20S(200S) 20(200) mol·m–3 sucrose  相似文献   

9.
Incorporation of 14CO2 in photosynthetic pigments of Chlorella pyrenoidosa   总被引:1,自引:1,他引:0  
Abscisic acid (ABA) caused a 7–8-fold increase in volume flow in excised bean root systems and this was coupled with an increase in 42K, 36Cl and 24Na flux into the xylem. The transport of 42K and 36Cl increased by a factor larger than the stimulation of volume flow, resulting in an increase in the concentration of those ions in the xylem exudate. Carbonyclcyanide-m-chlorophenyl hydrazone, on the other hand, eliminated ABA-stimulated 42K transport and caused a further inhibition of 42K flux, thus providing additional support for the proposition that ABA stimulation may involve an energised process of ion transport. ABA also increased the accumulation of 24Na and 36Cl in bean root tissue, but not that of 42K.  相似文献   

10.
R. Behl  W. Hartung 《Planta》1986,168(3):360-368
Epidermal peels of Valerianella locusta were acid-treated for 1 h at pH 3.9 to kill all cells other than guard cells. These guard-cell preparations were used to explore the steady-state one-way fluxes and the cytoplasmic and vacuolar contents of abscisic acid (ABA). The method of compartmental analysis has been applied. The intracellular ABA concentrations were surprisingly high. At an external pH of 5.8 the cytoplasm contained 1.28 mmol·dm-3 of ABA, twice of the amount which accumulated in the vacuoles (0.57 mmol·dm-3). The fluxes of ABA at the plasmalemma (oc=oc=0.43 fmol · cell –1 · h –1) were higher than those at the tonoplast (cv=vc=0.12 fmol · cell –1 · h –1). Moderate stress (0.1 and 0.3 mol·dm-3 sorbitol in the medium) caused a change in the kinetics of ABA movement. The rate constants of the fluxes from the cytoplasm into the vacuole (cv) and into the apoplast (co) were increased while the rate constant of the flux from the vacuoles into the cytoplasm (vc) was decreased. As a consequence the amount of ABA sequestered in the vacuole remained unchanged; the cytoplasmic ABA content, however, was reduced to only 20% of that found in the control treatments (no sorbitol in the medium). Under moderate stress, one Valerianella guard cell released rapidly about 0.36 fmol·cell-1 to its direct cell-wall space. This surprising result is discussed in regard to rapid stomatal closure under reduced water supply.Abbreviations ABA abscisic acid - FC fusicoccin  相似文献   

11.
Quantitative cell and organelle dynamics of the male gamete-producing lineage of Plumbago zeylanica were examined using serial transmission electron microscopic reconstruction at five stages of development from generative cell inception to sperm cell maturity. The founder population of generative cell organelles includes an average of 3.88 plastids, 54.9 mitochondria, and 3.7 vacuoles. During development the volume of the pollen grain increases from 6,200 μm3 in early microspores to 115,000 μm3 at anthesis, cell volume of the male germ lineage decreases more than 67% from 362.3 μm3 to 118.4 μm3. By the time the generative cell separates from the intine, plastid numbers increase by >600%, mitochondria by 250%, and vesicles by 43 times. A cellular projection elongates toward and establishes an association with the vegetative nucleus; this leading edge contains plastids and numerous mitochondria. When the generative cell completes its separation from the intine, organellar polarity is reversed and plastids migrate to the opposite pole of the cell. Cytoplasmic microtubules are common in association with cellular organelles. Plastids accumulate at the distal end of the cell as a linked mass, apparently adhered by lateral electron dense regions. Before division of the highly polarized generative cell, plastids decrease in number by 16%, whereas mitochondria increase by ∼90% and vacuoles increase by ∼140% from the prior stage. After mitosis, the resultant sperm cells differ in size and organelle content. The sperm cell associated with the vegetative nucleus (Svn) contains 62.7% of the cytoplasm volume, 87% of the mitochondria, 280.4 vesicles (79% of those in the generative cell), and 0.6% of the plastids. At maturity, the Svn mitochondria increase by 31% and the cell contains an average of 0.4 plastids, 158.9 vesicles, and 0.36 microbodies. The mature unassociated sperm (Sua) contains 39.8 mitochondria (up 3.3%), 24.3 plastids (down 31%), 91.1 vesicles (up 54.9%), and 3.18 microbodies. The small number of organelles initially in the generative cell, followed by their rapid multiplication in a shrinking cytoplasm suggests a highly competitive cytoplasmic environment that would tend to eliminate residual organellar heterogeneity. Cell and cytoplasmic volumes vary as a consequence of fluctuations in the number and size of large vesicles or vacuoles, as well as loss of cytoplasmic volume by (1) formation of “false cells” involving amitotic cytokinesis, (2) “pinching off” of cytoplasm, and (3) dehydration of pollen contents prior to anthesis.  相似文献   

12.
Data for the maximum carboxylation velocity of ribulose-1,5-biosphosphate carboxylase, Vm, and the maximum rate of whole-chain electron transport, Jm, were calculated according to a photosynthesis model from the CO2 response and the light response of CO2 uptake measured on ears of wheat (Triticum aestivum L. cv. Arkas), oat (Avena sativa L. cv. Lorenz), and barley (Hordeum vulgare L. cv. Aramir). The ratio Jm/Vm is lower in glumes of oat and awns of barley than it is in the bracts of wheat and in the lemmas and paleae of oat and barley. Light-microscopy studies revealed, in glumes and lemmas of wheat and in the lemmas of oat and barley, a second type of photosynthesizing cell which, in analogy to the Kranz anatomy of C4 plants, can be designated as a bundle-sheath cell. In wheat ears, the CO2-compensation point (in the absence of dissimilative respiration) is between those that are typical for C3 and C4 plants.A model of the CO2 uptake in C3–C4 intermediate plants proposed by Peisker (1986, Plant Cell Environ. 9, 627–635) is applied to recalculate the initial slopes of the A(pc) curves (net photosynthesis rate versus intercellular partial pressure of CO2) under the assumptions that the Jm/Vm ratio for all organs investigated equals the value found in glumes of oat and awns of barley, and that ribulose-1,5-bisphosphate carboxylase is redistributed from mesophyll to bundle-sheath cells. The results closely match the measured values. As a consequence, all bracts of wheat ears and the inner bracts of oat and barley ears are likely to represent a C3–C4 intermediate type, while glumes of oat and awns of barley represent the C3 type.Abbreviations A net photosynthesis rate (mol·m-2·s-1) - Jm maximum rate of whole-chain electron transport (mol·e-·m-2·s-1) - pc (bar) intercellular partial pressure of CO2 - PEP phosphoenolpyruvate - PPFD photosynthetic photon flux density (mol quanta·m-2·s-1) - RuBPCase ribulose bisphosphate carboxylase/oxygenase - RuBP ribulose bisphosphate - Vm maximum carboxylation velocity of RuBPCase (mol·m-2·s-1) - T* CO2 compensation point in the absence of dissimilative respiration (bar)  相似文献   

13.
Atripiex spongiosa was grown in hydroponic culture with additionof 0, 50, 200, 400 and 600 mM NaCl. Frozen, hydrated cells ofthe roots were analysed by X-ray micro-analysis. Comparisonswere made of meristematic cells at the root apex and vacuolatedcells 5.0 mm from the apex. High selectivitiy for K relativeto Na was found for the cytoplasm of meristematic cells andthere was little effect of increasing salinity on the ratiosNa/K, Cl/K, Na/P and CI/P. In the cell vacuoles of the cortex,selective uptake of K relative to Na also occurred, but to alesser extent than in the meristematic cells. Gradients werefound of decreasing ratio of Na/K from the epidermis to thestele. Measurements of chemical content of the roots and shoots ofthese plants showed that the ratio of Na/K was higher in theshoot than in the cortical cell vacuoles and higher again thanin the stele or meristematic cytoplasm. It is suggested thattransport of ions to the shoot of Atripiex spongiosa involvesselective exclusion of Na from the xylem parenchyma into thexylem, and that this may be general to other halophytes. Key words: Micro-analysis, X-ray, Cells, Atripiex spongiosa  相似文献   

14.
G. Hillmann  A. Ruthmann 《Planta》1982,155(2):124-132
After 5 h in 10-3M vinblastine the most obvious effects upon Vicia faba L. cells are seen in the spindle apparatus. These include the microtubules themselves as indicated by c-type metaphases and the pole regions of otherwise intact spindles, leading to multipolar anaphases and to telophases with more than two daughter nuclei. Vesicle transport may be undisturbed and new cell walls can be formed, although cases of disturbed cell plate and cell wall formation were noted occasionally, accompanied by myelinizations in phragmosomes. After 24 h in the same concentration of vinblastine, divisions are no longer observed and the plasma membranes are severely affected. They show extensive myelinizations, accumulations of lipids and dehiscence from the cell walls which are frequently thickened and irregularly formed. Of the other cellular membranes, the nuclear envelope and, more frequently, the tonoplast may be affected. Electron-dense deposits appear in the vacuoles. Comparable, though less severe, changes including multipolar anaphases and myelinizations result from treatment with lumicolchicine, but not with colchicine. Vinblastine leads to the appearance of filament bundles both in cytoplasm and karyoplasm, lumicolchicine to morphologically identical filaments in the cytoplasm alone. The nature of these filaments is unknown.Abbreviation VLB vincaleukoblastine  相似文献   

15.
The proposal that aluminium (Al) toxicity in plants is caused by either inhibition of Ca2+ influx or by displacement of Ca2+ from the cell wall, was examined. For this study the giant alga Chara corallina Klein ex Will. em. R.D. Wood was selected because it shows a similar sensitivity to Al as in roots of higher plants and, more importantly, it is possible to use the large single internodal cells to make accurate and unambiguous measurements of Ca2+ influx and Ca2+ binding in cell walls. Growth of Chara was inhibited by Al at concentrations comparable to those required to inhibit growth of roots, and with a similar speed of onset and pH dependence. At Al concentrations which inhibited growth, influx of calcium (Ca2+) was only slightly sensitive to Al. The maximum inhibition of Ca2+ influx at 0.1 mol·m–3 Al at pH 4.4 was less than 50%. At the same concentration, lanthanum (La3+) inhibited influx of Ca2+ by 90% but inhibition of growth was similar for both La3+ and Al. Removal of Ca2+ from the external solution did not inhibit growth for more than 8 h whereas inhibition of growth by Al was apparent after only 2.5 h. Ca2+ influx was more sensitive to Al when stimulated by addition of high concentrations of potassium (K+) or by action potentials generated by electrical stimulation. Other membrane-related activities such as sodium influx, rubidium influx and membrane potential difference and conductance, were not strongly affected by Al even at high concentrations. In isolated cell walls equilibrated in 0.5 mol·m–3 Ca2+ at pH 4.4, 0.1 mol·m–3 Al displaced more than 80% of the bound Ca2+ with a half-time of 25 min. From the poor correlation between inhibition of growth and reduction in Ca2+ influx, it was concluded that Al toxicity was not caused by limitation of the Ca2+ supply. Short-term changes in other membrane-related activities induced by Al also appeared to be too small to explain the toxicity. However the strong displacement, and probable replacement, of cell wall ca2+ by Al may be sufficient to disrupt normal cell development.Abbreviations CPW artificial pond water - PD potential difference The technical assistance of Dawn Verlin is gratefully acknowledged. This work was supported by the Australian Research Council.  相似文献   

16.
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   

17.
Temperature-dependent feedback inhibition of photosynthesis in peanut   总被引:7,自引:0,他引:7  
Arachis hypogaea L. is a tropical crop that is slow-growing at temperatures below 25°C. Unadapted CO2-assimilation rate (A) showed insufficient variation between 15 and 30°C in the short term (hours) to explain this marked reduction in growth. However, at longer periods (12 d), A was depressed as were growth rate and leafproduction rate. To examine the possible relationship between growth, A and sink demand plants were transferred from 30°C, which is near the optimum for growth, to a suboptimal temperature (19°C). In the first 2 d of cooling, A decreased by 50–70%, the stomata stayed open, and the intercellular CO2 concentration (ci) rose, i.e. the decrease in A of the cooled plants was the result of non-stomatal factors. Changes in dark respiration did not account for the decline in A.Clear evidence was obtained of sink control of A by independently manipulating the temperature of different leaves on the plant. Cooling (to 19°C) most of the plant (the sink) led to a 70% decline in A of the remaining leaves at 30°C after 3 d, whereas the converse treatments (30°C sink, 19°C source) resulted in small changes (17%). In plants at 19°C which were exposed to low CO2 concentration to prevent photosynthesis, A was not reduced when measured at normal CO2 concentrations, indicating that carbohydrate accumulation was responsible for the decline in A. Dry-matter build-up at suboptimal temperature was also consistent with end-product inhibition of photosynthesis.Abbreviations and symbols A (mol·m-2·s-1) rate of net CO2 assimilation - Ci (l·l-1) substomatal CO2 concentration - DW (g) dry weight - g (mol·m-2·s-1) stomatal conductance to diffusion of water vapour - PFD (mol·m-2·s-1) photon flux density  相似文献   

18.
The mechanism of zinc uptake in plants   总被引:1,自引:0,他引:1  
  相似文献   

19.
Hensel W 《Planta》1986,169(3):293-303
The development of the structural polarity of statocytes from cress roots (Lepidium sativum L.) was studied in a time- and stage-dependent manner. Outgrowing radicles had statocytes with abundant lipid droplets, sparsely developed endoplasmic reticulum (ER) and nuclei located at the proximal cell poles. During differentiation, coincidentally the lipid droplets disappeared, while rough ER increased in length. The ER was translocated into the distal cell pole to establish a complex of stacked ER. Microtubules occurred first at the distal cell edges. As a second step, ER was produced in the vicinity of the nucleus and was also translocated distally. By application of the antimicrotubular agents heavy water (90%), colchicine (10-4 mol·l-1) and triethyl lead chloride (20 mol·l-1), the involvement of microtubules in these events was studied. Triethyl lead chloride led to a complete cessation of differentiation; root-cap cells remained at a stage without polar arrangement of the ER. Colchicine affected the development of structural polarity slightly, as shown by a higher density of cortical ER cisternae. Heavy water inhibited the translocation of ER almost completely and yielded ER located also in the cell center. All anti-microtubular agents inhibited cell division and the differentiation of the distal cell layer of the dermatocalyptrogen into statocytes. It is hypothesized that microtubules serve as anchoring sites for microfilaments, which actually mediate the translocation of the ER. Hence, an intact system of microtubules and microfilaments is necessary for the expression of structural polarity.Abbreviations DC dermatocalyptrogen - ER endoplasmic reticulum - M meristem cell layer - MT microtubule - pI prospective story I - TrEl triethyl lead chloride  相似文献   

20.
J. J. Sauter  S. Kloth 《Planta》1986,168(3):377-380
The minimum radial translocation rate of sugars has been determined from the starchaccumulation rate for the wood rays of Populus x canadensis Moench robusta, and related to ultrastructural peculiarities of the cell walls to be passed. The minimum radial flux or flow of sugars through the tangential walls, the pit fields, and per plasmodesma was 80.7 pmol · cm-2 · s-1, 400 to 800 pmol · cm-2 · s-1, and 1.0 to 1.7 · 10-7 pmol · plasmodesma-1 · s-1, respectively. These values exclude a transmembrane flux mechanism and indicate that the radial translocation in this tissue must proceed via plasmodesmata. In the isolation cells of the ray center we found 39 plasmodesmata per m2 of pit field, 8.0 per m2 of tangential wall, and 1.98% of the wall occupied by plasmodesmata. Cells of the ray margins show plasmodesmata on only 1.16% of their tangential wall area and thus appear to be slightly inferior for radial translocation. As judged from both the observed plasmodesmatal frequencies and the translocation rates, the ray parenchyma cells are comparable to cells specialized in short-distance translocation.Abbreviations CCR contact-cell row - IC isolation cell - ICR isolation-cell row  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