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.
Osmotic water permeability of isolated vacuoles   总被引:5,自引:0,他引:5  
Morillon R  Lassalles JP 《Planta》1999,210(1):80-84
We measured the osmotic water permeability (P os) of vacuoles isolated from onion (Allium cepa L.), rape (Brassica napus L.), petunia (Petunia hybrida Hook.) and red beet (Beta vulgaris L.). For all the vacuolar types investigated, P os values were in the range 200–1000 μm s−1. The change in membrane surface area induced by an osmotic gradient was smaller than 2–6%. The vacuolar P os values for red beet and onion were reduced by 1 mM HgCl2, to 14% and 30% of the control values, respectively, but were partially restored to 51% and 76% by 5 mM β-mercaptoethanol. These results suggest that aquaporins were present in all the vacuoles tested. In HgCl2-treated onion vacuoles, the reduced P os (56 μm s−1) had a low activation energy (approx. 6 kJ mol−1), indicating that water permeation was still occurring mainly via aquaporins, and that the water permeability of the lipid part of the vacuolar membrane is probably very low. Received: 18 February 1999 / Accepted: 21 June 1999  相似文献   

2.
As part of a programme of comparative measurements of P d (diffusional water permeability) the RBCs (red blood cells) from an aquatic monotreme, platypus (Ornithorhynchus anatinus), and an aquatic reptile, saltwater crocodile (Crocodylus porosus) were studied. The mean diameter of platypus RBCs was estimated by light microscopy and found to be ~6.3 μm. P d was measured by using an Mn2+‐doping 1H NMR (nuclear magnetic resonance) technique. The P d (cm/s) values were relatively low: ~2.1×10?3 at 25°C, 2.5×10?3 at 30°C, 3.4×10?3 at 37°C and 4.5 at 42°C for the platypus RBCs and ~2.8×10?3 at 25°C, 3.2×10?3 at 30°C, 4.5×10?3 at 37°C and 5.7×10?3 at 42°C for the crocodile RBCs. In parallel with the low water permeability, the E a,d (activation energy of water diffusion) was relatively high, ~35 kJ/mol. These results suggest that “conventional” WCPs (water channel proteins), or AQPs (aquaporins), are probably absent from the plasma membranes of RBCs from both the platypus and the saltwater crocodile.  相似文献   

3.
The effects of amino acids (aa) and N-(diisopropyloxyphosphoryl)-amino acids (DIPP-aa) on cell membranes were investigated by evaluating water and methyl urea permeability. Permeability coefficients Pf and Ps were determined by standard osmotic methods for cells ofPisum sativum stem base epidermis after 20 min exposure to a 5 mM solution of each aa and DIPP-aa. The Pf value ofP. sativum epidermal cells (untreated controls) was 1.3 ± 0.4 × 10-3 μm s-1. Treat ments with the diisopropyl-oxyphosphoryl derivatives of three one charged and three polar amino acids (serine, threonine, asparagine, and aspartic acid) and unsubstituted (free) serine and threonine increased water permeability up to about two fold of the control value. Serine and threonine and their DIPP-derivatives increased methyl urea permeability (controls 1.03 ± 0.09 × 10-3 μm s-1) 30 to 80 percent Other amino acids and their DIPP-derivatives caused small or insignificant changes of water permeability. Only certain polar amino acids and their DIPP-derivatives increased the osmotic water and methyl urea permeation through the plasma membrane. The specificity of these molecules on plasma membranes suggests that the active amino acids (serine and threonine) and their DIPP-derivatives interact with charged membrane molecules. The relatively small changes in water and methyl urea permeability may indicate that the effective aa’s and their DIPP-derivatives interact with phospholipids rather than aquaporin. A concurring alteration of water channel proteins, however, cannot excluded.  相似文献   

4.
As part of a programme of comparative measurements of P d (diffusional water permeability) the RBCs (red blood cells) from dingo (Canis familiaris dingo) and greyhound dog (Canis familiaris) were studied. The morphologies of the dingo and greyhound RBCs [examined by light and SEM (scanning electron microscopy)] were found to be very similar, with regard to aspect ratio and size; the mean diameters were estimated to be the same (~7.2 μm) for both dingo and greyhound RBCs. The water diffusional permeability was monitored by using an Mn2+‐doping 1H NMR technique at 400 MHz. The P d (cm/s) values of dingo and greyhound RBCs were similar: 6.5×10?3 at 25°C, 7.5×10?3 at 30°C, 10×10?3 at 37°C and 11.5×10?3 at 42°C. The inhibitory effect of a mercury‐containing SH (sulfhydryl)‐modifying reagent PCMBS (p‐chloromercuribenzene sulfonate) was investigated. The maximal inhibition of dingo and greyhound RBCs was reached in 15–30 min at 37°C with 2 mmol/l PCMBS. The values of maximal inhibition were in the range 72–74% when measured at 25°C and 30°C, and ~66% at 37°C. The lowest value of P d (corresponding to the basal permeability to water) was ~2–3×10?3 cm/s in the temperature range 25–37°C. The E a,d (activation energy of water diffusion) was 25 kJ/mol for dingo RBC and 23 kJ/mol for greyhound RBCs. After incubation with PCMBS, the values of E a,d increased, reaching 46–48 kJ/mol in the condition of maximal inhibition of water exchange. The electrophoretograms of membrane polypeptides of the dingo and greyhound RBCs were compared and seen to be very similar. We postulate that the RBC parameters reported in the present study are characteristic of all canine species and, in particular in the two cases presented here, these parameters have not been changed by the peculiar Australian habitat over the millennia (as in the case of the dingo) or over shorter time periods, decades or centuries (as in the case of the domestic greyhound).  相似文献   

5.
The osmotic permeability coefficient (Pf) for water movement across Novikoff hepatoma cells was found to be 82 ± 3 (S.E.) · 10?5 cm · s?1 at 20°C. The corresponding diffusional permeability coefficient for 3HHO (Pd) was 97 ± 10 (S.E.) · 10?5 cm · s?1, therefore the ratio PfPd is close to unity. The apparent activation energy for water filtration was 10.4 ± 0.4 (S.E.) kcal · mol?1. This value is significantly greater than the activation energy for the self diffusion of water. The product of the hydraulic permeability coefficient and the viscosity coefficient for water was temperature-dependent. However, the product of the hydraulic permeability coefficient and the viscosity coefficient for membrane lipid did not vary with temperature. These data are interpreted as evidence for water movement across a lipid membrane barrier rather than through aqueous channels.  相似文献   

6.
The hydraulic conductance of the leaf lamina (Klamina) substantially constrains whole‐plant water transport, but little is known of its association with leaf structure and function. Klamina was measured for sun and shade leaves of six woody temperate species growing in moist soil, and tested for correlation with the prevailing leaf irradiance, and with 22 other leaf traits. Klamina varied from 7.40 × 10?5 kg m?2 s?1 MPa?1 for Acer saccharum shade leaves to 2.89 × 10?4 kg m?2 s?1 MPa?1 for Vitis labrusca sun leaves. Tree sun leaves had 15–67% higher Klamina than shade leaves. Klamina was co‐ordinated with traits associated with high water flux, including leaf irradiance, petiole hydraulic conductance, guard cell length, and stomatal pore area per lamina area. Klamina was also co‐ordinated with lamina thickness, water storage capacitance, 1/mesophyll water transfer resistance, and, in five of the six species, with lamina perimeter/area. However, for the six species, Klamina was independent of inter‐related leaf traits including leaf dry mass per area, density, modulus of elasticity, osmotic potential, and cuticular conductance. Klamina was thus co‐ordinated with structural and functional traits relating to liquid‐phase water transport and to maximum rates of gas exchange, but independent of other traits relating to drought tolerance and to aspects of carbon economy.  相似文献   

7.
A transference chamber was developed to measure the osmotic water permeability coefficient (Pos) in protoplasts 40 to 120 μm in diameter. The protoplast was held by a micropipette and submitted to a steep osmotic gradient created in the transference chamber. Pos was derived from the changes in protoplast dimensions, as measured using a light microscope. Permeabilities were in the range 1 to 1000 μm s−1 for the various types of protoplasts tested. The precision for Pos was ≤40%, and within this limit, no asymmetry in the water fluxes was observed. Measurements on protoplasts isolated from 2- to 5-d-old roots revealed a dramatic increase in Pos during root development. A shift in Pos from 10 to 500 μm s−1 occurred within less than 48 h. This phenomenon was found in maize (Zea mays), wheat (Triticum aestivum), and rape (Brassica napus) roots. These results show that early developmental processes modify water-transport properties of the plasma membrane, and that the transference chamber is adapted to the study of water-transport mechanisms in native membranes.  相似文献   

8.
Summary In order to assess the contribution of transcellular water flow to isosmotic fluid transport acrossNecturus gallbladder epithelium, we have measured the water permeability of the epithelial cell membranes using a nuclear magnetic resonance method. Spin-lattice (T 1) relaxation of water protons in samples of gallbladder tissue where the extracellular fluid contained 10 to 20mm Mn2+ showed two exponential components. The fraction of the total water population responsible for the slower of the two was 24±2%. Both the size of the slow component, and the fact that it disappeared when the epithelial layer was removed from the tissue, suggest that it was due to water efflux from the epithelial cells. The rate constant of efflux was estimated to be 15.6±1.0 sec1 which would be consistent with a diffusive membrane water permeabilityP d of 1.6×103 cm sec1 and an osmotic permeabilityP os of between 0.3×104 and 1.4×104 cm sec1 osmolar1. Using these data and a modified version of the standing-gradient model, we have reassessed the adequacy of a fluid transport theory based purely on transcellular osmotic water flow. We find that the model accounts satisfactorily for near-isosmotic fluid transport by the unilateral gallbladder preparation, but a substantial serosal diffusion barrier has to be included in order to account for the transport of fluid against opposing osmotic gradients.  相似文献   

9.
Given the increase in the incidence of insulin resistance, obesity, and type 2 diabetes in children and adolescents, it would be of paramount importance to assess quantitative indices of insulin secretion and action during a physiological perturbation, such as a meal or an oral glucose‐tolerance test (OGTT). A minimal model method is proposed to measure quantitative indices of insulin secretion and action in adolescents from an oral test. A 7 h, 21‐sample OGTT was performed in 11 adolescents. The C‐peptide minimal model was identified on C‐peptide and glucose data to quantify indices of β‐cell function: static φs and dynamic φd responsivity to glucose from which total responsivity φ was also measured. The glucose minimal model was identified on glucose and insulin data to estimate insulin sensitivity, SI, which was compared to a reference measure, SIref, provided by a tracer method. Disposition indices, which adjust insulin secretion for insulin action, were then calculated. Indices of β‐cell function were φs = 51.35 ± 8.89 × 10?9min?1, φd = 1,392 ± 258 × 10?9, and φ = 82.09 ± 17.70 × 10?9min?1. Insulin sensitivity was SI = 14.19 ± 2.73 × 10?4, not significantly different from SIref = 14.96 ± 3.04 × 10?4 dl/kg·min per µU/ml, and well correlated: r = 0.98, P < 0.0001, thus indicating that SI can be accurately measured from an oral test. Disposition indices were DIs = 1,040 ± 201 × 10?14 dl/kg/min2 per pmol/l, DId = 33,178 ± 10,720 × 10?14 dl/kg/min per pmol/l, DI = 1,844 ± 522 × 10?14 dl/kg/min2 per pmol/l. Virtually the same minimal model assessment was obtained with a reduced 3 h, 9‐sample protocol. OGTT interpreted with C‐peptide and glucose minimal model has the potential to provide novel insight regarding the regulation of glucose metabolism in adolescents, and to evaluate the effect of obesity and interventions such as diet and exercise.  相似文献   

10.
Water and solute relations of young roots of Phaseolus coccineus have been measured using the root pressure probe. Biphasic root pressure relaxations were obtained when roots were treated with solutions containing different osmotic test solutes. From the relaxations, the hydraulic conductivity (Lpr), the permeability coefficient (Psr), and the reflection coefficient (σsr) of the roots could be evaluated. Lpr was 1.8 to 8.4 . 10?8 m . s?1 . MPa?1 and Psr (in 10?10 m . s?1): methanol, 27–62; ethanol, 44–73; urea, 5–11; mannitol, 1.5; KCl, 7.1–9.2; NaCl, 2.1; NaNO3, 3.7. The hydraulic conductivity was similar when using osmotic and hydrostatic pressure gradients as driving forces. The hydraulic conductivity of individual root cortex cells (Lp) was by two orders of magnitude larger than Lpr (Lp = 0.3 to 4.7 . 10?6 m . s?1 . MPa?1) which indicated a predominant cell-to-cell rather than an apoplasmic transport of water in the Phaseolus root. Except for distances shorter than 20 mm from the root apex, the hydraulic resistance of the roots was limited by the radial movement of water across the root cylinder and not by the hydraulic resistance within the xylem. Reflection coefficients were low: methanol: 0.16–0.34; ethanol: 0.15–0.47; urea: 0.41–0.51; mannitol: 0.68; KCl: 0.43–0.54; NaCl: 0.59; NaNO3: 0.54. The transport coefficients (Lpr, Psr, σsr) have been critically examined for influences of unstirred layers and active transport. The low σsr suggests that the common treatment of the root as a rather perfect osmometer (σsr = 1) analogous to plant cells should be treated cautiously. The reasons for the low σsr and the possible implications of the absolute values of the transport parameters for the absorption of water and nutrients are discussed.  相似文献   

11.
Brush border membrane vesicles, BBMV, from eel intestinal cells or kidney proximal tubule cells were prepared in a low osmolarity cellobiose buffer. The osmotic water permeability coefficient P f for eel vesicles was not affected by pCMBS and was measured at 1.6 × 10−3 cm sec−1 at 23°C, a value lower than 3.6 × 10−3 cm sec−1 exhibited by the kidney vesicles and similar to published values for lipid bilayers. An activation energy E a of 14.7 Kcal mol−1 for water transport was obtained for eel intestine, contrasting with 4.8 Kcal mol−1 determined for rabbit kidney proximal tubule vesicles using the same method of analysis. The high value of E a , as well as the low P f for the eel intestine is compatible with the absence of water channels in these membrane vesicles and is consistent with the view that water permeates by dissolution and diffusion in the membrane. Further, the initial transient observed in the osmotic response of kidney vesicles, which is presumed to reflect the inhibition of water channels by membrane stress, could not be observed in the eel intestinal vesicles. The P f dependence on the tonicity of the osmotic shock, described for kidney vesicles and related to the dissipation of pressure and stress at low tonicity shocks, was not seen with eel vesicles. These results indicate that the membranes from two volume transporter epithelia have different mechanisms of water permeation. Presumably the functional water channels observed in kidney vesicles are not present in eel intestine vesicles. The elastic modulus of the membrane was estimated by analysis of swelling kinetics of eel vesicles following hypotonic shock. The value obtained, 0.79 × 10−3 N cm−1, compares favorably with the corresponding value, 0.87 × 10−3 N cm−1, estimated from measurements at osmotic equilibrium. Received: 28 January 1999/Revised: 15 June 1999  相似文献   

12.
Zoeae of Paralithodes camtschatica were positively phototactic to white light intensities above 1 × 1013 q cm?2 s?1. Negative phototaxis occurred at low (1 × 1012 q cm?2 s?1), but not high intensities (2.2 × 1016q cm?2 s?1). Phototactic response was directly related to light intensity. Zoeae also responded to red, green and blue light. Zoeae were negatively geotactic, but geotaxis was dominated by phototaxis. Horizontal swimming speed of stage 1 zoeae <4 d old was 2.4 ± 0.1 (SE) cms?1 and decreased to 1.7 ± 0.1 cm s?1 in older zoeae (P <0.01). Horizontal swimming speed of stage 2 zoeae was not significantly different from ≥4 d old stage 1 zoeae. Vertical swimming speed, 1.6 ± 0.1 cm s?1, and sinking rate, 0.7 ± 0.1 cm s?1, did not change with ontogeny. King crab zoeae were positively rheotactic and maintained position in horizontal currents less than 1.4 cm s?1. Starvation reduced swimming and sinking rates and phototactic response.  相似文献   

13.
The permeability of secondary E. granulosus cysts to [14C]mebendazole was studied. The cysts were obtained by transplanting secondary cysts raised in mice into rats. The permeability to [14C]mebendazole was established by two different experiments: uptake and washout of the drug. The cyst wall permeability to [14C]mebendazole was found to be 1·33 × 10?4 cm s?1, which is of the same order as the diffusion permeability coefficient to water (1·88 × 10?4 cm s?1, Rotunno, Kammerer, Perez Esandi & Cereijido, 1974).The drug readily permeates through the cyst wall and experimental data suggest that it moves across the barrier by simple diffusion.  相似文献   

14.
1. The ability of hyporheic sediments to exchange water and retain ammonium and phosphate in the Riera Major stream ,North-East Spain, under different discharge conditions was measured by conducting short-term nutrient and chloride additions. 2. The mean exchange coefficients from free-flowing water to the storage zone (k1) and vice versa (k2) were 0.82 × 10–4 s??1 and 7 × 10??3 s??1, respectively. The ratio of storage zone cross-sectional area to stream cross-sectional area (AS/A) averaged 2.8 × 10–2 and was negatively correlated with discharge (r = –0.85, d.f. = 13, P < 0.001). 3. The percentage of hyporheic zone water which came from surface water varied as a function of discharge and hyporheic depth, ranging between 33% and 95% at 25 cm depth, and between 78% and 100% at 10 cm depth. 4. The nutrient retention efficiency in the hyporheic zone at 10 cm depth measured as uptake length (Swh) was less than 3.3 cm for ammonium and 37 cm for phosphate. Higher nutrient retentions were measured in the sediments at 10 cm depth than at 25 cm, indicating that near-surface sediments were involved more actively in phosphate retention than the deeper hyporheic sediments. The lack of ammonium at any depth of the hyporheic zone showed that ammonium was very rapidly taken up in the surfacial sediments.  相似文献   

15.
The diffusional water permeability (Pd) in various gut structures of the fleshfly, Sarcophaga bullata, was measured using tritiated water. Water Pd in the larval hindgut was 3.91 × 10?6 cm/sec, whereas in the adult hindgut it was 4.4 × 10?4 cm/sec. The presence of cuticle in various parts of the gut apparently controls the water permeabilities of these structures. Furthermore, the water permeability of the cuticle may be correlated with the mechanism for the production of a hyperosmotic excretion in the hindgut.  相似文献   

16.
It is demonstrated that cyanobacteria (both azotrophic and non‐azotrophic) contain heme b oxidoreductases that can convert chlorite to chloride and molecular oxygen (incorrectly denominated chlorite ‘dismutase’, Cld). Beside the water‐splitting manganese complex of photosystem II, this metalloenzyme is the second known enzyme that catalyses the formation of a covalent oxygen–oxygen bond. All cyanobacterial Clds have a truncated N‐terminus and are dimeric (i.e. clade 2) proteins. As model protein, Cld from Cyanothece sp. PCC7425 (CCld) was recombinantly produced in Escherichia coli and shown to efficiently degrade chlorite with an activity optimum at pH 5.0 [kcat 1144 ± 23.8 s?1, KM 162 ± 10.0 μM, catalytic efficiency (7.1 ± 0.6) × 106 M?1 s?1]. The resting ferric high‐spin axially symmetric heme enzyme has a standard reduction potential of the Fe(III)/Fe(II) couple of ?126 ± 1.9 mV at pH 7.0. Cyanide mediates the formation of a low‐spin complex with kon = (1.6 ± 0.1) × 105 M?1 s?1 and koff = 1.4 ± 2.9 s?1 (KD ~ 8.6 μM). Both, thermal and chemical unfolding follows a non‐two‐state unfolding pathway with the first transition being related to the release of the prosthetic group. The obtained data are discussed with respect to known structure–function relationships of Clds. We ask for the physiological substrate and putative function of these O2‐producing proteins in (nitrogen‐fixing) cyanobacteria.  相似文献   

17.
18.
The effects of nitrogen (N) nutrition on growth, N uptake and leaf osmotic potential of rice plants (Oryza sativa L. ev. IR 36) during simulated water stress were determined. Twenty-one-day-old seedlings in high (28.6 × 10 ?4M) and low (7.14 × 10 4M) N levels were exposed to decreased nutrient solution water potentials by addition of polyethylene glycol 6000. The roots were separated from the solution by a semi-permeable membrane. Nutrient solution water potential was ?0.6 × 105 Pa and was lowered stepwise to ?1 × 105, ?2 × 105, ?4 × 105 and ?6 × 105 Pa at 2-day intervals. Plant height, leaf area and shoot dry weight of high and low nitrogen plants were reduced by lower osmotic potentials of the root medium. Osmotic stress caused greater shoot growth reduction in high N than in low N plants. Stressed and unstressed plants in 7.14 × 104M N had more root dry matter than the corresponding plants in 28.6 × 104M N. Dawn leaf water potential of stressed plants was 1 × 105 to 5.5 × 105 Pa lower than nutrient solution water potential. Nitrogen-deficient water-stressed plants, however, maintained higher dawn leaf water potential than high nitrogen water-stressed plants. It is suggested that this was due to higher root-to-shoot ratios of N deficient plants. The osmotic potentials of leaves at full turgor for control plants were about 1.3 × 105 Pa higher in 7.14 × 10?4M than in 28.6 × 10?4M N and osmotic adjustment of 2.6 × 105 and 4.3 × 105 Pa was obtained in low and high N plants, respectively. The nitrogen status of plants, therefore, affected the ability of the rice plant to adjust osmotically during water stress. Plant water stress decreased transpiration and total N content in shoots of both N treatments. Reduced shoot growth as a result of water stress caused the decrease in amount of water transpired. Transpiration and N uptake were significantly correlated. Our results show that nitrogen content is reduced in water-stressed plants by the integrated effects of plant water stress per se on accumulation of dry matter and transpiring leaf area as well as the often cited changes in soil physical properties of a drying root medium.  相似文献   

19.
Annett Hertel  Ernst Steudle 《Planta》1997,202(3):324-335
Using the cell pressure probe, the effects of temperature on hydraulic conductivity (Lp; osmotic water permeability), solute permeability (permeability coefficient, Ps), and reflection coefficients (σs) were measured on internodes of Chara corallina, Klein ex Willd., em R.D.W.. For the first time, complete sets of transport coefficients were obtained in the range between 10 and 35 °C which provided evidence about pathways of water and solutes as they move across the plasma membrane (water channel and bilayer arrays). Test solutes used to check for the selectivity of water channels were monohydric alcohols of different molecular size and shape (ethanol, n-propanol, iso-propanol, and tert-butanol) and heavy water (HDO). Within the limits of accuracy, Q10 values for Lp and for the diffusive water permeability (Pd) were identical (Q10 for Lp = 1.29 ± 0.17 (± SD; n = 15 cells) and Q10 for Pd = 1.25 ± 0.16 (n = 5 cells)). The Q10 values were equivalent to activation energies of Ea = 16.8 ± 6.4 and 16.6 ± 10.0 kJ · mol−1, respectively, which is similar to that of self-diffusion or of viscous flow of water. The Q10 values and activation energies for Ps of the alcohols were significantly larger (ethanol: Q10 = 1.68 ± 0.16, Ea = 37.1 ± 5.9 kJ · mol−1; n-propanol: Q10 =  1.75 ± 0.40, Ea = 43.1 ± 15.3 kJ · mol−1; iso-propanol: Q10 = 2.12 ± 0.42, Ea =  52.2 ± 14.6 kJ · mol−1; tert-butanol: Q10 = 2.13 ± 0.56, Ea = 51.6 ± 17.1 kJ · mol−1; ±SD; n = 5 to 6 cells). Effects of temperature on reflection coefficients were most pronounced. With increasing temperature, σs values of the alcohols decreased and those of HDO increased. The data indicate that water and solutes use different pathways when crossing the membrane. Ordinary and isotopic water use water channels and the other test solutes use the bilayer array (composite transport model of membrane). Changes in σs values with temperature were found to be a sensitive measure for the open/closed state of water channels. The decrease of σs with temperature was theoretically predicted from the temperature dependence of Ps and Lp. Differences between predicted and measured values of σs allowed estimation of the bypass flow (slippage) of solutes through water channels which did not completely exclude test solutes. The permeability of channels depended on the structure and size of test solutes. It is concluded that water channels are much less selective than is usually thought. Since water channels represent single-file or no-pass pores, solutes drag along considerable amounts of water as they diffuse across channels. This results in low overall values of σs. The σs of HDO was extremely low. Its response to temperature was opposite to that for the σs of the alcohols. This suggested a stronger effect of temperature on the hydraulic (osmotic) than on the diffusive water flow across individual water channels, i.e. a differential sensitivity of different mechanisms to temperature. Received: 10 October 1996 / Accepted: 2 December 1996  相似文献   

20.
Responses of leaf and shoot hydraulic conductance to light quality were examined on shoots of silver birch (Betula pendula), cut from lower (‘shade position’) and upper thirds of the crowns (‘sun position’) of trees growing in a natural temperate forest stand. Hydraulic conductances of leaf blades (Klb), petioles (KP) and branches (i.e. leafless stem; KB) were determined using a high pressure flow meter in steady state mode. The shoots were exposed to photosynthetic photon flux density of 200–250 µmol m?2 s?1 using white, blue or red light. Klb depended significantly on both light quality and canopy position (P < 0.001), KB on canopy position (P < 0.001) and exposure time (P = 0.014), and none of the three factors had effect on KP. The highest values of Klb were recorded under the blue light (3.63 and 3.13 × 10?4 kg m?2 MPa?1 s?1 for the sun and shade leaves, respectively), intermediate values under white light (3.37 and 2.46 × 10?4 kg m?2 MPa?1 s?1, respectively) and lowest values under red light (2.83 and 2.02 × 10?4 kg m?2 MPa?1 s?1, respectively). Light quality has an important impact on leaf hydraulic properties, independently of light intensity or of total light energy, and the specific light receptors involved in this response require identification. Given that natural canopy shade depletes blue and red light, Klb may be decreased both by reduced fluence and shifts in light spectra, indicating the need for studies of the natural heterogeneity of Klb within and under canopies, and its impacts on gas exchange.  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