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.
Water-relation parameters (cell turgor pressure [P], volumetric elastic modulus [epsilon] and hydraulic conductivity [Lp]) of individual leaf epidermal cells of Tradescantia virginiana have been determined with the pressure-probe technique. Turgor was 4.5 +/- 2.1 [41] bar (mean +/- sd; in brackets the number of cells) and ranged from 0.9 to 9.6 bar. By vacuum infiltration with nutrient solution, it was raised to 7.5 +/- 1.5 [5] bar (range: 5.3-8.8 bar). There was a large variability in the absolute value of epsilon of individual cells. epsilon ranged from 40 to 360 bar; mean +/- sd: 135 +/- 83 bar; n = 50 cells. epsilon values of individual cells seemed to be rather independent of changes in cell turgor. A critical assessment of the errors incurred in determining epsilon by the technique is included. The half-times of water exchange of individual cells ranged from 1 to 35 seconds, which gave values of 0.2 to 11 x 10(-6) centimeters per second per bar for Lp (mean +/- sd: 3.1 +/- 2.3 x 10(-6) centimeters per second per bar; n = 39 cells). The large range in Lp and epsilon is believed to be due to the difficulties in determining the effective surface area of water exchange of the cells. Lp is not influenced by active salt pumping driven by respiration energy inasmuch as it was not altered by 0.1 millimolar KCN. The temperature dependence of Lp (T((1/2))) was measured for the first time in individual higher-plant cells. Lp increased by a factor of 2 to 4, when the temperature was increased by 10 C. The activation energy of water exchange was found to be between 50 and 186 kilojoules per mole. Within the large range of variation it was found that T((1/2)), Lp, and epsilon did not change under various experimental conditions (intact and excised tissue, water content and turgidity, age, etc.). Similar results were obtained for the epidermal cells of Tradescantia andersoniana. The measurements suggest that the entire epidermis would respond very rapidly (i.e. with a half-time of 1 to 30 s) to a demand for water from the stomata.  相似文献   

2.
Using the automated cell pressure probe, small and highly reproducible hydrostatic pressure clamp (PC) and pressure relaxation (PR) tests (typically, applied step change in pressure = 0.02 MPa and overall change in volume = 30 pL, respectively) were applied to individual Tradescantia virginiana epidermal cells to determine both exosmotic and endosmotic hydraulic conductivity (LpOUT and LpIN, respectively). Within-cell reproducibility of measured hydraulic parameters depended on the method used, with the PR method giving a lower average coefficient of variation (15.2%, 5.8%, and 19.0% for half-time, cell volume [Vo], and hydraulic conductivity [Lp], respectively) than the PC method (25.4%, 22.0%, and 24.2%, respectively). Vo as determined from PC and PR tests was 1.1 to 2.7 nL and in the range of optically estimated Vo values of 1.5 to 4.9 nL. For the same cell, Vo and Lp estimates were significantly lower (about 15% and 30%, respectively) when determined by PC compared with PR. Both methods, however, showed significantly higher LpOUT than LpIN (LpOUT/LpIN ≅ 1.20). Because these results were obtained using small and reversible hydrostatically driven flows in the same cell, the 20% outward biased polarity of water transport is most likely not due to artifacts associated with unstirred layers or to direct effects of externally applied osmotica on the membrane, as has been suggested in previous studies. The rapid reversibility of applied flow direction, particularly for the PR method, and the lack of a clear increase in LpOUT/LpIN over a wide range of Lp values suggest that the observed polarity is an intrinsic biophysical property of the intact membrane/protein complex.The conductivity of membranes to water (hydraulic conductivity [Lp]) is an important property of the cells of all organisms, and whether plant cell membranes exhibit a polarity in this property has been debated for a number of decades (Dainty and Hope, 1959; Steudle, 1993). Most early evidence for polarity was based on transcellular osmotic experiments using giant algal cells in the Characeae, in which the relative areas of cell membrane exposed to conditions of osmotic inflow (endosmosis) or outflow (exosmosis) could be varied and, hence, Lp for both directions determined (Tazawa and Shimmen, 2001). Interpretation of these experiments is complicated by unstirred layer (USL) effects (Dainty, 1963), but even after accounting for these, it was concluded that inflow Lp (LpIN) was higher than outflow Lp (LpOUT) in these cells, with LpOUT/LpIN of about 0.65 (Dainty, 1963). When using osmotic driving forces in algal cells, LpOUT/LpIN values of between 0.5 and 0.91 have been reported in many studies (Steudle and Zimmermann, 1974; Steudle and Tyerman, 1983; Tazawa et al., 1996), and the same direction of polarity was also reported using osmotic driving forces in whole roots of maize (Zea mays; Steudle et al., 1987). When applying hydrostatic driving forces in algal cells using the pressure probe (Steudle, 1993), which is less influenced by USL effects (Steudle et al., 1980), LpOUT/LpIN has been closer to 1 (0.83–1; Steudle and Zimmermann, 1974; Steudle and Tyerman, 1983). However, in higher plant cells, an analysis of the data presented by Steudle et al. (1980, 1982) and Tomos et al. (1981) indicates the opposite polarity, with LpOUT/LpIN averaging from 1.2 to 1.4. Moore and Cosgrove (1991) used two contrasting hydrostatic methods to measure Lp in sugarcane (Saccharum spp.) stem cells: (1) the most commonly used pressure relaxation (PR) method, in which cell turgor pressure (Pcell) changes during the measurement, and (2) the more technically demanding pressure clamp (PC) method, in which Pcell is maintained constant. Consistent with other studies in higher plant cells, Moore and Cosgrove (1991) reported average LpOUT/LpIN from 1.15 (PC) to 1.65 (PR). Using the PR method in epidermal cells of barley (Hordeum vulgare), Fricke (2000) reported only a modest LpOUT/LpIN (based on reported half-time [T1/2]) of 1.08. In view of the contribution of proteins (e.g. aquaporins) to overall membrane Lp, Tyerman et al. (2002) suggested that polarity may result either from asymmetry in the pores themselves or from an active regulation of the conductive state of the pores in response to the experimental conditions that cause inflow or outflow. Either of these mechanisms may explain the wide range of values reported in the literature for LpOUT/LpIN. Cosgrove and Steudle (1981) reported that a substantial (6-fold) and rapid (within 20 s) reduction in Lp could occur in the same cell, and in hindsight, this presumably reflected the influence of aquaporins. Cosgrove and Steudle (1981) did not consider the lower Lp as indicative of the Lp in situ, and Wan et al. (2004) reported that a reduction in Lp was associated with perturbations to Pcell on the order of 0.1 MPa. Hence, if measured membrane Lp itself can exhibit substantial changes over relatively short periods of time in the same cell, then further study of systematic differences between LpOUT and LpIN will require a robust hydrostatic methodology (PC or PR) that can reversibly and reproducibly apply small perturbations in pressure (P) to individual cells over short periods of time.For the PR method, a T1/2 of water exchange is measured by fitting an exponential curve to the observed decay in Pcell over time following a step change in volume, and membrane Lp can be calculated if cell surface area (A), cell volume (Vo), and volumetric elastic modulus (ε) are known (Steudle, 1993). In practice, A and Vo are typically calculated from optical measurements of individual cell dimensions or estimates using average values, and ε is calculated based on Vo and an empirical change in pressure (dP) to change in volume (dV) relation for each cell (Steudle, 1993; Tomos and Leigh, 1999). In the PC method, first developed by Wendler and Zimmermann (1982), Vo (and, given reasonable assumptions about cell geometry, A) is estimated without the need for optical measurements, and Lp can be measured without the need to determine dP/dV or ε. However, this method is technically more demanding because it requires precise P control as well as a continuous record of the volume flow of water across the cell membrane (as measured by changes in the position of the cell solution/oil meniscus within the glass capillary over time) and has rarely been used (Wendler and Zimmermann, 1982, 1985; Cosgrove et al., 1987; Moore and Cosgrove, 1991; Zhang and Tyerman, 1991; Murphy and Smith, 1998). Since volume (V) is continuously changing over time, this approach may also be influenced by the hydraulic conductance of the capillary tip (Kh) used to make the measurements as well as surface tension effects due to the progressive changes in capillary diameter with meniscus position, and these influences have not been quantitatively addressed.Automation of the pressure probe operation, particularly automatic tracking of the meniscus location in the glass microcapillary tip, would address many of the above-mentioned issues, and to date, several attempts have been made to monitor the meniscus location using electrical resistance (Hüsken et al., 1978) or hardware-based image analysis (Cosgrove and Durachko, 1986; Murphy and Smith, 1998). Recently, Wong et al. (2009) redesigned the automated cell pressure probe (ACPP), originally proposed by Cosgrove and Durachko (1986), using a software-based meniscus detection system and a precise pressure control system. In the new ACPP system, both the position of the meniscus and oil pressure (Poil) are recorded frequently (typically at 10 Hz), and Poil is controlled with a resolution of ±0.002 MPa. We have combined the ACPP with a new technique to reproducibly fabricate microcapillary tips of known hydraulic properties (Wada et al., 2011) in order to correct for Kh and surface tension effects in both PC and PR estimates of the water relations parameters of Tradescantia virginiana epidermal cells and have determined the relation of LpOUT to LpIN in these cells.  相似文献   

3.
Cell membrane potentials have been measured both in epidermalstrips and intact leaf sections of Tradescantia virginiana andCommelina communis, and in epidermal cells over green and overalbino mesophyll cells of T. albiflora var. albovittata. Membranepotentials (cell) in strips were considerably lower than thosein intact sections and were insensitive to light and to theabsence or presence of calcium. Their response to external cationlevels was indifferent to ionic species. However, in intactleaf sections incubated with calcium present, membrane potentialsresponded to K+ levels but not to Na+. were more negative thancells in epidermal strips, and responded to changes in illumination. Long-term recordings of cell and vacuolar K+ levels in T. virginianaduring stomatal closure suggest that the fluctuations of cellwere unrelated to K+ movement (which we could not detect) andthus probably to stomatal movement as well. Turgor pressures measured in epidermal cells of intact leafsections of T. virginiana were found to be of the same magnitudeas those previously reported for epidermal strips. It is concludedthat epidermal cells maintain their solute contents during strippingwithout the involvement of an electrophysiological transportsystem. With the possible exception of lateral subsidiary cells,there was no evidence suggesting that ordinary epidermal cellsare capable of osmotic adjustment even when additional KCI wassupplied in the osmoticum. Absolute turgor levels in intactleaf sections kept at constant external KCI were unrelated tosteady state cell.  相似文献   

4.
Cuticular resistance to water vapour diffusion between the substomatalcavity and the sensing psychrometer junction is a problem uniqueto leaf hygrometry. This resistance is not encountered in soilor solution hygrometry. The cuticular resistance may introduceerror in the measurement of leaf water potential. Using in situleaf hygrometers, we studied the effect of abrading the cuticleof Citrus jambhiri Lushington leaves, to reduce the diffusiveresistance. Field measurements of psychrometer water potentialwere compared with Scholander pressure chamber values for adjacentleaves. Different treatments were compared by sealing pairsof psychrometers on either side of the midrib. The time forwater vapour equilibration between the leaf and the psychrometerchamber was greater than 5 h for no abrasion. For abraded leaves,the true water potential value was obtained within an hour.After equilibration, psychrometer values compared favourablywith pressure chamber values for adjacent leaves (r > 0.97).Measured water potential for unabraded leaves did not correlatewell with corresponding pressure chamber measurements. Scanning electron micrographs indicated that the damage causedby abrading leaves for 60 s using carborundum powder (60 µmdiameter) was surface localized, with numerous scratchings ofthe leaf cuticle. The coarse abrasion treatment (aluminium oxide,75 µm diameter) resulted in fewer but larger cavitiesin the epidermis, which may explain the observed variabilityin the corresponding psychrometric measurements. Key words: Leaf water potential, Cuticular resistance, Leaf abrasion, Thermocouple psychrometer  相似文献   

5.
Water Pathways in Leaves of Hedera helix L. and Tradescantia virginiana L.   总被引:2,自引:0,他引:2  
Hydraulic conductances of leaf tissues of Hedera helix and Tradescantiavirginiana leaves were measured. It was found that water couldflow most easily through the veins, but that the cell wallsof at least the ventral epidermis were more efficient at resupplyingwater lost from the epidermal tissue than was the mesophyllat rehydrating itself. Vein and bundle-sheath extensions, whichare characteristic of mesomorphic leaves (e.g. T. virginiana),seem to be important in maintaining a close hydraulic connectionbetween the epidermis and the vascular tissue. In leaves notcontaining vein and bundle-sheath extensions, typically xeromorphicleaves (e.g. H. helix), there is not such a close connectionbetween the epidermis and vascular tissue. This was shown inexperiments involving the sudden application of a reduced pressurepotential to either the epidermis or the other tissues of leaves,and the measurement of transient stomatal opening.  相似文献   

6.
Liu  M.Z.  Jiang  G.M.  Li  Y.G.  Gao  L.M.  Niu  S.L.  Cui  H.X.  Ding  L. 《Photosynthetica》2003,41(3):393-398
Gas exchange, photochemical efficiency, and leaf water potential (l) of Salix matsudana (non-indigenous species), S. microstachya and S. gordejevii (indigenous species) were studied in Hunshandak Sandland, China. l of all the three species decreased from 06:00 to 12:00, and increased afterwards. S. matsudana showed higher values of l than others. Net photosynthetic rate (P N) and stomatal conductance (g s) of S. matsudana were the lowest among all, with the maximum P N at 10:00 being 75% of that of S. gordejevii. Compared with the indigenous species, the non-indigenous S. matsudana had also lower transpiration rate (E) and water use efficiency (WUE). The values of Fv/Fm in all the species were lower from 06:00 to 14:00 than those after 14:00, indicating an obvious depression in photochemical efficiency of photosystem 2 in both non-indigenous and native species. However, it was much more depressed in S. matsudana, the non-indigenous tree. P N was positively correlated to g s and negatively related to l. The relationship between g s and vapour pressure difference (VPD) was exponential, while negative linear correlation was found between g s and l.  相似文献   

7.
SHERIFF  D. W. 《Annals of botany》1982,50(4):535-548
The hydraulic conductances of leaves of a species which exhibitsstomatal responses to humidity (Nicotiana glauca) are significantlylower than the conductances in a species which does not exhibitsuch responses (Tradescantia virginiana). This difference couldat least partly account for their difference in stomatal responseto humidity. In both species, the hydraulic conductance betweenthe leaf bulk and its epidermis is much lower than the conductancein any other part of the pathway. The apparently conflictingresults, reported in recent literature, on the hydraulic conductancesand water pathways in leaves are reinterpreted, and shown tobe due to misinterpretation of results. The recently publishedcriticisms of a technique used to measure hydraulic conductivityare commented on and refuted. An examination of the factors that influence the water potentialat the sites of evaporation from the inner walls of the epidermisnear stomatal pores showed that the water potential at thesesites is lower than the bulk epidermal water potential. Thewater potential at these sites changes in a complex way as stomatalaperture changes. As it is reduced the ratio of: ‘waterpotential at sites of evaporation on the inner walls of theepidermis near stomatal pores/bulk leaf water potential‘increases. The positive feedback effect of this phenomenon,which tends to keep stomatal water potential constant as thestomata close and therefore enhances closure, and two other‘passive’ positive feedback effects on the waterpotential at sites of evaporation near stomata that have beenreported in the literature are briefly discussed. Nicotiana glauca (Grah.), Tradescantia virginiana (L.), sub-stomatal cavities, peristomatal evaporation, stomata, humidity response, leaf hydraulic conductance, water potential  相似文献   

8.
Luteolin, kaempferol, quercetin, caffeic acid and ferulic acidwere identified in acid-hydrolyzed epidermal strips of Tradescantiavirginiana using HPLC and spectrophotometry. The amount of flavonoidswas much smaller than that of cinnamic acid derivatives. Morethan 80% of the flavonoids were found in methanol extracts ofepidermal strips. Caffeic acid was found in both methanol extractsand the residues in nearly equal amounts, while more than 80%of the ferulic acid was found in the residues after methanolextraction. These data suggest that most of the ferulic acidand part of the caffeic acid bind to macromolecules as estersin the cell wall and that flavonoids are localized mainly inthe cytoplasm. The localization of esters of hydroxycinnamicacids in cell walls was ascertained by fluorometric analysis.These phenolic compounds were oxidized by H2O2 (0.025–1mM) in epidermal and guard cells and the oxidation was inhibitedby KCN and NaN3: luteolin glycosides were less sensitive toH2O2 than quercetin and kaempferol glycosides in flavonoids.Ferulic acid esters were more sensitive to H2O2 than caffeicacid esters in hydroxycinnamic acid derivatives. On the basisof these data, the physiological significance of the oxidationof phenolic compounds by H2O2 is discussed. (Received October 9, 1987; Accepted February 3, 1988)  相似文献   

9.
When the apertures of stomata in an epidermal strip changedthere was found to be a simultaneous change in the unbound watercontent of the tissue. This occurred even when only the guardcells were living. It was concluded that the change in unboundwater occurred in the guard cells, and possibly the guard cellwalls.  相似文献   

10.
The physiological basis underlying differences in sensitivity of different aged leaves to water stress was investigated in Fragaria virginiana Duchesne. Differential susceptibility of only older leaves to water stress in the field during summer months appeared related to gradients in leaf osmotic potential within the plant and by an age dependency in the ability of leaves to adjust osmotically when challenged by periodic water deficits. Under greenhouse conditions, older leaves senesced invariably during an imposed water stress while control leaves of comparable age and stressed younger leaves remained green. Osmotic potentials of intermediate aged and younger leaves became approximately 1 to 2 bars lower after a single cycle of imposed stress and up to 10 bars lower after two cycles of stress. Pronounced gradients in leaf osmotic potential within individual whole plants were observed following two cycles of water stress that were significantly different from control values. Osmotic adjustment was dependent on leaf age with the greatest capacity for adjustment in the intermediate aged leaves. Loss of osmotic adjustment was rapid upon rewatering with a half-life of 4 days. An irreversible component of adjustment was observed, amounting to about 10% (or 2 bars) of the maximally adjusted state. This irreversible component could be accounted for in part by significant changes in cell size and other anatomical alterations in the leaf that affect cellular osmotic volume, and, hence, cellular water relations.  相似文献   

11.
以生长于浑善达克沙地上的C3植物白榆(Ulmus pumila)、C4植物沙米(Agriophyllum pungens)和CAM植物钝叶瓦松(Orostachys malacophyllus)3种不同光合途径植物为材料,测定了它们生长期叶片的光合气体交换参数、叶绿素荧光参数和水势,探讨它们对生长环境的生理响应特征.结果表明,白榆和沙米的净光合速率、气孔导度均高于钝叶瓦松,特别是在夏季高温(>40℃)和强光照(>2 100 μmol·m-2·s-1)条件下表现得更加明显.白榆和沙米的光合速率、叶片水势都发生了严重的午休现象,其白天光合速率的降低主要是由于气孔关闭造成的.钝叶瓦松的叶片水势在3种植物中最高,但是白天的光合速率很低;其Fv/Fm值在14:00最低,一天中此时光系统II受伤害最大;CAM物种瓦松的碳固定仅发生在夜间.研究发现,C3植物白榆和C4植物沙米比CAM植物钝叶瓦松对热和高光照有着更强的忍耐力,瓦松固定碳主要发生在生长最快的阶段;CAM植物瓦松为了能够在夏季强光和高温条件下生存,它必须进行高强度的呼吸,仅在早晨和夜间进行碳固定.  相似文献   

12.
An instrument is described which permits the non-destructivemeasurement of the mean pressure potential (turgor)of leaf laminacells. Calibration shows that the instrument gives a voltageoutput which is linearly related to mean pressure potentialof living leaf cells as determined with a pressure chamber.Measurements may be made very quickly in the field or controlledenvironment with a resolution of at least 50 k Pa.  相似文献   

13.
WOLFE  DAVID W. 《Annals of botany》1991,67(3):205-212
Two chilling-sensitive (Phaseolus vulgaris L., Zea mays L.)and two chilling-tolerant (Pisum sativum L., Spinacia oleraceaL.) species were raised in growth chambers under warm (28/18°Cday/night cycle) and cool (18/12°C) temperature regimes.Growth analysis techniques were used to evaluate leaf area andbiomass partitioning during early autotrophic growth. Plantsacclimated to both temperatures were measured for leaf gas exchangeand water potential (  相似文献   

14.
Growth, Turgor, Water Potential, and Young's Modulus in Pea Internodes   总被引:1,自引:0,他引:1  
The relations between longitudinal growth, Young's modulus, turgor, water potential, and tissue tensions have been studied on growing internodes of etiolated pea seedlings in an attempt to apply some physical concepts to the growth of a well-known plant material. The modulus has been determined by the resonance frequency method and expressed as Etissue It increases nearly proportional to the turgor pressure and is at water saturation more than 50 times higher than at plasmolysis. Etissue is higher in the epidermis than in the ground parenchyma. Indoleacetic acid causes a decrease in Etissue Other properties have been studied on intact and split segments of internodes in solutions of graded mannitol additions. — The following tentative picture of the normal course of the growth has been obtained. Auxin induces growth both in the periphery (epidermis) and in the central core (parenchyma) under a decrease in Etissue This is followed by an increase of Etissue which is independent of auxin but depending upon the turgor pressure. It is assumed to involve internal structural changes of the cell walls of the type of creep. The rapid growth takes place in a dynamic system with a low water potential despite favourable water conditions. Epidermis and parenchyma grow equally rapid without tissue tensions. — Such can be produced artificially by splitting of segments and water uptake. The parenchyma thereby loses its sensitivity to auxin. This is the background of the split stem test for auxin. — Etissue increases when growth is slowing down, probably owing to both synthesis of wall substance and structural changes within the wall. The cells attain a more static condition with Etissue higher in epidermis than in parenchyma. This leads to the normal tissue tensions. — The result agrees with growth according to the multi-net-principle. The cause of the low water potential and low turgor is discussed with reference to the dynamic nature of both growth and water transport and a probably low matric potential of the streaming water. The decrease in Etissue following auxin addition is small but is the net difference between an auxin-induced decrease and an increase through the assumed creep.  相似文献   

15.
Illuminated leaf discs of Vicia faba were brought into equilibrium with a series of mannitol solutions. The width of stomatal aperture and the osmotic potential of guard cells and epidermal cells were determined. It was found that the maximal aperture was obtained when epidermal cells were at about incipient plasmolysis and that any increase in their turgor pressure brought about a decrease in stomatal aperture. These findings emphasize the importance of epidermal cells in determining the width of the stomatal pore.  相似文献   

16.
Abstract. The interrelationships of leaf diffusive conductance, tap root cell turgor pressure and the diameter of the tap root of sugarbeet were studied. The study was conducted on well-watered plants growing in pots under artificial light in the glasshouse. In a typical experiment, on illumination (400 μmol m−2 s−1) leaf conductance increased from 0.6 to 7.4 mm s−1. Cell turgor pressure in the tap root decreased from 0.8 MPa to 0.45 MPa and the root diameter (9.0 cm) contracted by 145μm. Removal of light resulted in the reversal of each of the above parameters to their previous values. Quantitively similar results were obtained when sugar beet plants were uprooted and the response of each of the parameters was measured. The sequence of events however was different. On stimulation by light, changes in leaf diffusive conductance preceded the turgor and root diameter changes (which were simultaneous) by some 15–20min. In contrast, on uprooting the simultaneous changes in root turgor pressure and diameter preceded the changes in leaf conductance. The lag times between changes in diffusive conductance and turgor pressure in the root were between 20 and 30 min.
Tap root turgor pressure and diameter correlated strongly and permitted the calculation of an apparent whole root volumetric elastic modules (55–63 MPa). The small changes in tissue volume relative to the transpiration rate suggest that the tap root is not a significant source of transpirational water during the day.  相似文献   

17.
Frensch J  Hsiao TC 《Plant physiology》1994,104(1):247-254
Transient responses of cell turgor (P) and root elongation to changes in water potential were measured in maize (Zea mays L.) to evaluate mechanisms of adaptation to water stress. Changes of water potential were induced by exposing roots to solutions of KCl and mannitol (osmotic pressure about 0.3 MPa). Prior to a treatment, root elongation was about 1.2 mm h-1 and P was about 0.67 MPa across the cortex of the expansion zone (3-10 mm behind the root tip). Upon addition of an osmoticum, P decreased rapidly and growth stopped completely at pressure below approximately 0.6 MPa, which indicated that the yield threshold (Ytrans,1) was just below the initial turgor. Turgor recovered partly within the next 30 min and reached a new steady value at about 0.53 MPa. The root continued to elongate as soon as P rose above a new threshold (Ytrans,2) of about 0.45 MPa. The time between Ytrans,1 and Ytrans,2 was about 10 min. During this transition turgor gradients of as much as 0.15 MPa were measured across the cortex. They resulted from a faster rate of turgor recovery of cells deeper inside the tissue compared with cells near the root periphery. Presumably, the phloem was the source of the compounds for the osmotic adjustment. Turgor recovery was restricted to the expansion zone, as was confirmed by measurements of pressure kinetics in mature root tissue. Withdrawal of the osmoticum caused an enormous transient increase of elongation, which was related to only a small initial increase of P. Throughout the experiment, the relationship between root elongation rate and turgor was nonlinear. Consequently, when Y were calculated from steady-state conditions of P and root elongation before and after the osmotic treatment, Yss was only 0.21 MPa and significantly smaller compared with the values obtained from direct measurements (0.42-0.64 MPa). Thus, we strongly emphasize the need for measurements of short-term responses of elongation and turgor to determine cell wall mechanics appropriately. Our results indicate that the rate of solute flow into the growth zone could become rate-limiting for cell expansion under conditions of mild water stress.  相似文献   

18.
Abstract: An increasing body of evidence suggests that in plants, as in vertebrates, biologically active natriuretic peptide (NP) hormones play an important role in the regulation of the osmotic and ionic balance. The evidence includes isolation and immunoaffinity purification of biologically active natriuretic peptide analogues (irPNP) from ivy that promoted stomatal opening and specifically, rapidly and transiently increased cGMP levels in root conductive tissue. In this study we demonstrate that I125-rat atrial natriuretic peptide (rANP) binds to plasma membranes from leaf and stem tissue of Tradescantia multiflora and importantly, both unlabelled rANP and irPNP can competitively displace that binding. In addition, tissue section autoradiography reveals specific in situ binding of I125-rANP to leaf and stem tissue. The findings are consistent with the presence of a biologically active NP system in plants and suggest that NPs signal through a dedicated receptor system.  相似文献   

19.
Pressure and Solute Potentials in Stomatal Cells of Tradescantia virginiana   总被引:4,自引:0,他引:4  
The postulated mechanical advantage of subsidiary cells overguard cells has been estimated using leaves of Tradescantiavirginiana. The turgor pressures of subsidiary cells were adjusted to bezero or maximal by plasmolytic treatments, and the resultingstomatal apertures were measured. The ‘mechanical advantage’was calculated from two mathematical models which define itas an ‘antagonism ratio’. The discussion deals withmethods of preparing the tissue, the validity of the plasmolytictreatments, and the function of the antagonism ratio in relationto the Spannungsphase  相似文献   

20.
The tonoplast of Tradescantia virginiana L. was prepared from leaf cells and then solubilized with deoxycholate (DOC) and n-octyl-beta-D-glucoside (n-OG). Three major polypeptides (68, 60, 16 kDa) and several other minor components were isolated. These polypeptides were reconstituted in soybean phospholipids (asolectin). The H(+) pump activity was investigated with the reconstituted system as well as with the tonoplast. In both cases, the quinacrine-fluorescence quenching was observed in the presence of ATP-Mg(2+), indicating the H(+) pumping. The H(+) pump activity was inhibited by gramicidin D, a channel-forming ionophore, and by KNO(3), an inhibitor specific to tonoplast-type (V-type) H(+)-ATPase.  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