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.
E. Steudle  J. S. Boyer 《Planta》1985,164(2):189-200
Hydraulic resistances to water flow have been determined in the cortex of hypocotyls of growing seedlings of soybean (Glycine max L. Merr. cv. Wayne). Data at the cell level (hydraulic conductivity, Lp; half-time of water exchange, T 1/2; elastic modulus, ; diffusivity for the cell-to-cell pathway, D c) were obtained by the pressure probe, diffusivities for the tissue (D t) by sorption experiments and the hydraulic conductivity of the entire cortex (Lpr) by a new pressure-perfusion technique. For cortical cells in the elongating and mature regions of the hypocotyls T 1/2=0.4–15.1 s, Lp=0.2·10-5–10.0·10-5 cm s-1 bar-1 and D c=0.1·10-6–5.5·10-6 cm2 s-1. Sorption kinetics yielded a tissue diffusivity D t=0.2·10-6–0.8·10-6 cm2 s-1. The sorption kinetics include both cell-wall and cell-to-cell pathways for water transport. By comparing D c and D t, it was concluded that during swelling or shrinking of the tissue and during growth a substantial amount of water moves from cell to cell. The pressure-perfusion technique imposed hydrostatic gradients across the cortex either by manipulating the hydrostatic pressure in the xylem of hypocotyl segments or by forcing water from outside into the xylem. In segments with intact cuticle, the hydraulic conductance of the radial path (Lpr) was a function of the rate of water flow and also of flow direction. In segments without cuticle, Lpr was large (Lpr=2·10-5–20·10-5 cm s-1 bar-1) and exceeded the corticla cell Lp. The results of the pressure-perfusion experiments are not compatible with a cell-to-cell transport and can only the explained by a preferred apoplasmic water movement. A tentative explanation for the differences found in the different types of experiments is that during hydrostatic perfusion the apoplasmic path dominates because of the high hydraulic conductivity of the cell wall or a preferred water movement by film flow in the intercellular space system. For shrinking and swelling experiments and during growth, the films are small and the cell-to-cell path dominates. This could lead to larger gradients in water potential in the tissue than expected from Lpr. It is suggested that the reason for the preference of the cell-to-cell path during swelling and growth is that the solute contribution to the driving force in the apoplast is small, and tensions normally present in the wall prevent sufficiently thick water films from forming. The solute contribution is not very effective because the reflection coefficient of the cell-wall material should be very small for small solutes. The results demonstrate that in plant tissues the relative magnitude of cell-wall versus cell-to-cell transport could dependent on the physical nature of the driving forces (hydrostatic, osmotic) involved.Abbreviations and symbols D c diffusivity of the cell-to-cell pathway - D t diffusivity of the tissue - radial flow rate per cm2 of segment surface - Lp hydraulic conductivity of plasma-membrane - Lpr radial hydraulic conductance of the cortex - T 1/2 half-time of water exchange between cell and surroundings - volumetric elastic modulus  相似文献   

2.
Summary Tradescantia virginiana L. plants were cultivated under contrasting conditions of temperature, humidity, light quality and intensity, and nutrient status in order to investigate the effect of growth conditions on the water relations parameters of the leaf epidermal cells. Turgor pressure (P), volumetric elastic modulus (), half-time of water potential equilibration (T 1/2), hydraulic conductivity (L p ) were measured with the miniaturized pressure probe in single cells of the upper and lower epidermis of leaves. Turgor differed (range: 0.1 bar to 7.2 bar) between treatments with lowest values under warm and humid conditions and additional supply of fertilizer, and highest values under conditions of low air humidity and low nutrient supply. The volumetric elastic modulus changed by 2 orders of magnitude (range: 3.0 bar to 350 bar, 158 cells), but was only affected by the treatments, in as much as it was dependent on turgor. The turgor dependence of , measured on intact leaves of T. virginiana, was similar to that for cells of the isolated (peeled) lower epidermis, where as a function of turgor was linear over the whole range of turgors. This result has implications for the discussion of pressure/volume curves as measured by the pressure bomb where changes in bulk leaf are frequently discussed as adaptations to certain treatments. The measurements of the hydraulic conductivity indicate that this parameter varies between treatments (range of means: 2.4×10-6 cm s-1 bar-1 to 13.4×10-6 cm s-1 bar-1). There was a negative correlation for L p in cells of intact leaves as a function of turgor which was altered by the growing conditions. However, a correlation with turgor could not be found for cells from isolated epidermis or cells from a uniform population of plants. The large variation in L p from cell to cell observed in the present and in previous studies was accounted for in a study of 100 cells from a uniform population of plants by the propagation of measurement errors in calculating L p . The results suggest that in T. virginiana cellular water relations are changed mainly by the turgor dependence of .  相似文献   

3.
Water relation parameters of leaf cells of the aquatic plant Elodea densa have been measured using the pressure probe. For cells in both the upper and lower epidermis it was found that the elastic modulus () and the hydraulic conductivity (Lp) were dependent on cell turgor (P). Lp was (7.8±5.5)·10-7 cm s-1 bar-1 (mean±SD; n=22 cells) for P>4 bar in cells of the upper epidermis and was increasing by a factor of up to three for P0 bar. No polarity of water movement or concentration dependence of Lp was observed. For cells of the lower epidermis the Lp-values were similar and the hydraulic conductivity also showed a similar dependence on turgor. No wall ingrowth or wall labyrinths (as in transfer cells) could be found in the cells of the lower epidermis. The elastic modulus () of cells of the upper epidermis could be measured over the whole pressure range (P=0–7 bar) by changing the osmotic pressure of the medium. increased linearly with increasing turgor and ranged between 10 and 150 bar. For cells of the lower epidermis the dependence of on P was similar, although the pressure dependence could not be measured on single cells. The Lp-values are compared with literature data obtained for Elodea by a nuclear magnetic resonance (NMR)-technique. The dependence of Lp on P is discussed in terms of pressure dependent structural changes of the cell membranes and interactions between solute and water transport.Abbreviations P cell turgor pressure - Lp hydraulic conductivity - volumetric elastic modulus - T 1/2 half-time of water exchange of individual cell  相似文献   

4.
The turgor-homeostat model of assimilate efflux from coats of developing seed of Phaseolus vulgaris L. was further characterised. The turgor pressure (P), the volumetric elastic modulus () and hydraulic conductivity (Lp) of the seed coat cells responsible for assimilate efflux and cotyledon storage parenchyma cells were determined with a pressure probe. In addition, turgor of the seed coat and cotyledons was estimated by measuring the osmolalities of symplastic and apoplastic fluids extracted by centrifugation. Osmolality of symplastic and apoplastic saps collected from the seed coat declined significantly over the period of seed development from a cotyledon water content of 80% to 50%. However, the difference in osmolalities of the apoplastic and symplastic saps remained relatively constant. For cotyledons, osmolality of the apoplastic sap exhibited a significant decline during seed development, while the osmolality of symplastic sap did not change significantly. Hence cotyledon P increased as the water content dropped from 80% to 50%. For both detached and attached empty seed coats, a small decrease (ca. 40mOsmol·kg–1) in the osmolality of the bathing solution, led to a rapid increase in P of cells involved in assimilate efflux (efflux cells) by about 0.07 MPa. Thereafter, cell P exhibited a rapid decline to the original value within some 20–30 min. When P of the efflux cells was reduced by increasing the osmolality of the bathing solution, P exhibited a comparable rate of recovery for attached empty seed coats but there was no P recovery to its original value in the case of detached seed coats. In contrast, the cotyledon storage parenchyma cells did not exhibit P regulation when the osmolality of the bathing solution was changed. The observations that the efflux cells of P. vulgaris seed coats can rapidly adjust their P homeostatically in response to small changes in apoplastic osmolality are consistent with the operation of a turgor-homeostat mechanism. The volumetric elastic modulus () of the seed coat efflux cells exhibited a mean value of 7.3±0.8 MPa at P=0.15 MPa and was found to be linearly dependent on cell P. The e of the cotyledon storage parenchyma cells was estimated to be 6.1±1.0 MPa at P=0.41 MPa. Hydraulic conductivity (Lp) of the seed coat cells and the cotyledon cells was (8.2±1.5) × 10–8m·s–1·MPa–1and (12.8±1.0) × 10–8 m·s–1·MPa–1, respectively. The relatively high , i.e., low elasticity, for the seed coat cell walls would ensure that small changes in water potential of the seed apoplast will be reflected in large changes in cell P. The high Lp values for both the seed coat and the cotyledon cells is consistent with the rapid changes in P in response to changes in water potential of the seed apoplast.Abbreviations LYCH Lucifer Yellow CH - volumetric elastic modulus - Lp hydraulic conductivity - P turgor pressure - osmotic pressure - t1/2 half-time for water exchange The investigation was supported by funds from the Australian Research Council. We are grateful to Louise Hetherington for competent technical assistance and to Kevin Stokes for raising the plant material.  相似文献   

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

6.
All of the cells of the upper (adaxial) epidermis of the leaves ofOxalis carnosa are transformed into large bladders, while in the lower epidermis the bladder cells are interrupted by “normal” cells with stomata. The epidermal bladders contain a high concentration of free oxalic acid (pH approx. 1). Water-relations parameters of these epidermal bladder cells have been determined using the pressure probe. Original cell turgor (P0) of the closely packed bladders of theupper epidermis was P0=0.7 to 2.9 bar ( \(\overline {P_0 } = 1.7 \pm 0.5 bar\) ; mean±SD;N=25 cells) and lower than that in the club-shaped bladders of thelower epidermis (P0=1.3 to 3.7 bar; \(\overline {P_0 } = 2.5 \pm 0.7 bar\) ;N=25 cells). Large differences in the elastic modulus (ε) and the hydraulic conductivity (Lp) of the two different types of cells were observed. For the lower epidermal bladders, ε=18 to 166 bar and was similar to that of other higher plant cells. Also, for these cells it was found that ε was increasing with both, cell turgor and cell volume. By contrast, ε of the cells of the upper epidermis was by one order of magnitude smaller (ε=1.9 to 17.0 bar) and no dependence of ε on cell volume could be detected. The Lp values of the cell membranes were also different (lower epidermis: \(\overline {Lp} = (2.3 \pm 1.6) \cdot 10^{ - 5} cm s^{ - 1} bar^{ - 1}\) ; upper epidermis: \(\overline {Lp} = (3.8 \pm 2.4) \cdot 10^{ - 6} cm s^{ - 1} bar^{ - 1}\) ). These differences seem to be too large to be caused by errors in determining the exchange area for water (A) between cells and adjacent tissue. The half-times of water exchange between bladders and leaf (T1/2) were, on average, somewhat longer for the upper than for the lower epidermis (lower epidermis: T1/2=7 to 38 s; upper epidermis: T1/2=22 to 213 s), but the differences in the T1/2 values were not as distinct as for ε and Lp. This is because of the compensatory effects of ε, Lp and the different ratios of volume to exchange area. Since the bladders make up about 75% of the entire volume of the leaf, it is assumed that the rate of response of the leaf to changes in the water potential should be similar to that of the bladder cells. The results are discussed in terms of a possible function of the bladders in the leaf.  相似文献   

7.
Summary The effect of 1-alkanols upon the main phase-transition temperature of phospholipid vesicle membranes between gel and liquid-crystalline phases was not a simple monotonic function of alkanol concentration. For instance, 1-decanol decreased the transition temperature at low concentrations, but increased it at high concentrations, displaying a minimal temperature. This concentration-induced biphasic effect cannot be explained by the van't Hoff model on the effect of impurities upon the freezing point. To explain this nonlinear response, a theory is presented which treats the effect of 1-alkanols (or any additives) on the transition temperature of phospholipid membranes in a three-component mixture. By fitting the experimental data to the theory, the enthalpy of the phase transition H * and the interaction energy, AB * between the additive and phospholipid molecules may be estimated. The theory predicts that when AB * >2 (where AB * = AB,/RT o,T o being the transition temperature of phospholipid), both maximum and maximum transition temperatures should exist. When AB * = 2, only one inflection point exists. When AB * < 2, neither maximum nor minimum exists. The alkanol concentration at which the transition temperature is minimum (X min) depends on the AB * value: the larger the AB * values, the smaller theX min. When AB * is large enough,X min values become so small that the plot T vs.X shows positive T in almost all alkanol concentrations. The interaction energy between 1-alkanols and phospholipid molecules increased with the increase in the carbon chain-length of 1-alkanols. In the case of the dipalmitcylphosphatidylcholine vesicle membrane, the carbon chain-length of 1-alkanols that caused predominantly positive T was about 12.  相似文献   

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

9.
One cultivar each of spring wheat (Triticum aestivum L. cv. Arkas), oat (Avena sativa L. cv. Lorenz), and barley (Hordeum vulgare L. cv. Aramir) was chosen in order to study the relative contributions of individual bracts to the gas exchange of whole ears. The distribution and frequency of the stomata on the bracts were examined. Gas exchange was measured at normal atmospheric CO2 (330 bar) and at high CO2 (2000 bar) on intact ears and on ears from which glumes or lemmas and pleae (wheat and oat) or awns (barley) had been removed.The relative contribution to the gas exchange of the whole organ is highest for the awns of barley ears. In wheat, the contribution of the glumes is slightly higher than that of the inner bracts before anthesis. Two weeks after anthesis the inner bracts contribute more than the glumes. This tendency of increasing importance of the inner bracts is also found in oat ears, but the relative amount of CO2 uptake by the glumes is higher than in wheat. These changes during ontogeny result from the better supply of light to the inner bracts caused by opening of the ears' structures during grain filling, which in part compensates for the decreasing photosynthetic capacity.The ratio of the photosynthesis rate at high CO2 to that at normal CO2 is lower for the glumes of oat and for the awns of barley than for the other bracts.Abbreviations A330, A2000 net photosynthesis rate, A330 at normal atmospheric CO2 (330 bar), A2000 at high CO2 (2000 bar) - PPFD photosynthetic photon flux density - pc intercellular partial pressure of CO2  相似文献   

10.
During ageing of the short-lived pollen grains of Cucurbita pepo L., water loss was examined in relation to viability using biophysical (1H-nuclear magnetic resonance, NMR) and cytological methods (fluorochromatic reaction test, freezefracture and scanning electron microscopy). A semi-logarithmic representation of the pollen weight loss demonstrated the complexity of the dehydration process. A the study of proton loss using 1H-NMR indicated that two major releases water of had taken place, each with different flux rates. Pulse 1H-NMR experiments showed the occurrene of non-exponential signal decay as a function of time, indicating the existence of different fractions of water in a pollen grain sample. These fractions leave the pollen grain at different times during pollen dehydration, and one of them (that of the so-called vital water) can be related to pollen viability. The quantity of protons giving a signal during pulse 1H-NMR experiments was very low when the pollen grains were judged to be dead according to the fluorochromatic test. Freeze-fracture replicas of these dead pollen grains (less than 25% water content) showed that the plasma membrane had become detached from the intine surface; this ultrastructural feature might therefore be involved in the loss of pollen viability.Abbreviations A initial amplitude of the NMR signal - A2 quantity of water charcterized by T2-2 - A5 quantity of water characterized by T2–5 - FCR fluorochromatic reaction - NMR nuclear magnetic resonance - T2 transverse relaxation time - T2-2 T2 measured with 2 ms between each pulse of radiofrequency - T2–5 T2 measured with 5 ms between each pulse of radiofrequency  相似文献   

11.
Summary Using a pressure probe, turgor pressure was directly determined in leaf-mesophyll cells and the giant epidermal bladder cells of stems, petioles and leaves of the halophilic plant Mesembryanthemum crystallinum. Experimental plants were grown under non-saline conditions. They displayed the photosynthetic characteristics typical of C3-plants when 10 weeks old and performed weak CAM when 16 weeks old. In 10 week old plants, the turgor pressure (P) of bladder cells of stems was 0.30 MPa; of bladder cells of petioles 0.19 MPa, and of bladder cells of leaves 0.04 MPa. In bladder cells from leaves of 16 week old plants, marked changes in turgor pressure were observed during day/night cycles. Maximum turgor occurred at noon and was paralleled by a decrease in the osmotic pressure of the bladder cell sap. Similar changes in the cell water relations were observed in plants in which traspirational water loss was prevented by high ambient relative humidity. Turgor pressure of mesophyll cells also increased during day-time showing macimum values in the early morning. No such changes in turgor pressure and osmotic pressure were observed in bladder and mesophyll cells of the 10 week old plants not showing the diurnal acid fluctuation typical of CAMAbbreviations CAM crassulacean acid metabolism - V volume of the cells (mm3) - P turgor pressure (MPa) - volumetric elastic modulus (MPa) - i osmotic pressure of the cell sap (MPa) - T 1/2 half-time of water exchange (s) - Lp hydraulic conductivity of the cell membrane (m·s-1·MPa-1) - A surface area of cells (mm2) - P pressure changes (MPa) - V volume changes (mm3) - nocturanal nighttime - diurnal daytime  相似文献   

12.
Measurement of the light response of photosynthetic CO2 uptake is often used as an implement in ecophysiological studies. A method is described to calculate photosynthetic parameters, such as the maximum rate of whole electron transport and dissimilative respiration in the light, from the light response of CO2 uptake. Examples of the light-response curves of flag leaves and ears of wheat (Triticum aestivum cv. ARKAS) are shown.Abbreviations and symbols A net photosynthesis rate - D 1 rate of dissimilative respiration occurring in the light - f loss factor - I incident PPFD - I effective absorbed PPFD - J rate of whole electron transport - J m maximum rate of whole electron transport - p c intercellular CO2 partial pressure - PPFD photosynthetic photon flux density - q effectivity factor for the use of light (electrons/quanta) - absorption coefficient - I * CO2 compensation point in the absence of dissimilative respiration (bar) - II conversion factor for calculation of CO2 uptake from the rate of whole electron transport - convexity factor Gas-exchange rates relate to the projective area and are given in mol·m-2·s-1. Electron-transport rates are given in mol electrons·m-2·s-1; PPFD is given in mol quanta·m-2·s-1.  相似文献   

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

14.
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   

15.
The apparent photosynthetic Km (CO2) of air-grown Dunaliella salina is 2 M as measured both by the filtering centrifugation technique and by O2 electrode. These cells are capable of accumulating inorganic carbon (Cinorg) up to 20 times its concentration in the medium. It is suggested that air-grown Dunaliella cells are able to concentrate CO2 within the cell. Analysis of the efflux of Cinorg from cells previously loaded with H14CO 3 - demonstrated the existence of an internal pool which has an half-time of depletion of 2.5–7 min depending on the conditions of the experiment. This finding indicates that the internal Cinorg pool is not readily exchangeable with the external medium. Furthermore, the influence of the presence or absence of unlabelled Cinorg in the medium during the efflux experiment on the half-time observed indicate that efflux of Cinorg is not a simple diffusion process but is rather carrier-mediated.Abbreviation Cinorg inorganic carbon  相似文献   

16.
Zusammenfassung Es wurden analysenreine Proben der Romanowsky-Farbstoffe Eosin Y, Erythrosin B und Tetrachlorfluoreszein hergestellt.Im DC der Farbstoffproben konnten keine Verunreinigungen nachgewiesen werden. Die Absorptionsspektren der Farbstoffdianionen in wäßriger alkalischer Lösung und der Farbstoffsäuren in 95%igem Ethanol wurden bei sehr kleinen Farbstoffkonzentrationen gemessen und der molare Extinktionskoeffizient der längstwelligen Absorptionsbande der monomeren Farbstoffspezies bestimmt (Tabelle 1). Die Extinktionskoeffizienten können zur Standardisierung von Farbstoffproben verwendet werden. Die Absorptionsspektren von Eosin Y hängen in wäßriger Lösung von der Farbstoffkonzentration ab. Aus der Konzentrationsabhängigkeit wurden mit einem neuen, sehr empfindlichen Verfahren zwei Assoziationsgleichgewichte ermittelt. Bereits in sehr verdünnter Lösung bilden sich Dimere, bei erhöhter Konzentration Tetramere, Die Dissoziationskonstante der DimerenD in MonomereM beträgt bei pH=12, 293K:K 21=2,9 × 10–5 M; der TetramerenQ in DimereD:K 42=2,4 × 10–3 M. Aus den gemessenen Spektren von Eosinlösungen verschiedener Konzentration, pH=12, und den GleichgewichtskonstantenK 21,K 42 haben wir die Spektren der reinen Monomeren, Dimeren und Tetrameren bestimmt.M hat eine langwellige Absorptionsbande: , M =1,03 x 105 M-1 cm-1;D eine Bande: , D =1,74 x 105 M-1 cm-1;Q zwei Banden: , , Q1=1,65 x 105, Q2=1,96 x 105 M-1 cm-1. Das Absorptionsspektrum der Dimeren wird quantenmechanisch interpretiert.
Romanowsky dyes and Romanowsky-Giemsa effect. 2. Eosin Y, Erythrosin B, tetrachlorofluorescein, Spectroscopic characterization of pure dyes, association of Eosin Y
Summary Analytically pure smaples of the Romanowsky dyes eosin y, erythrosin b and tetrachlorofluorescein are prepared. DC of the dye samples shows no contaminations. We measured the absorption spectra of the dye dianions in alkaline aqueous solution and of the dye acids in 95% ethanol at very low dye concentrations. The molar extinction coefficients of the long wavelength absorption of the monomeric dye species are determined (Table 1). The extinction coefficients may be used for standardisation of dye samples. The absorption spectra of eosin y in aqueous solution are dependend on concentration. Using a new very sensitive method it was possible to identify two association equilibria from the concentration dependency of the spectra. Dimers are formed even in very dilute solutions, at higher concentrations tetramers. The dissociation constant of the dimersD in monomersM at 293 K, pH=12, isK 21=2,9×10–5 M; of the tetramersQ in dimersDK 42=2,4×10–3 M. From the experimental spectra of eosin solutions at various concentrations, pH=12, and the equilibrium constantsK 21,K 42 the absorption spectra of the pure monomers, dimers and tetramers are calculated. M has one long wavelength absorption band, , M =1,03 x 105 M-1 cm-1;D also one absorption band, , D =1,74 x 105 M-1 cm-1;Q two absorption bands, , , Q1=1,65 x 105, Q2=1,96 x 105 M-1 cm-1. The absorption spectrum of the dimers is discussed by quantum mechanics.
  相似文献   

17.
Summary Methylated lysines (N -mono-methylated, N -di-methylated and N -tri-methylated) have been identified after derivatization with orthophthaldialdehyde (OPA) by using pre-column and post-column derivatization techniques.Also the N -acetylated lysine and N -formylated lysine have been identified by OPA post-column derivatization techniques but only in free form due to their instability under acidic conditions which are used for protein hydrolysis.Additionally, all the modified amino acids mentioned above have been derivatized with DABITC/PITC, an Edman reagent, and identified as DABTH-derivatives on thin-layer polyamide sheets.  相似文献   

18.
Ian E. Woodrow  Keith A. Mott 《Planta》1993,191(4):421-432
A model of the C 3 photosynthetic system is developed which describes the sensitivity of the steadystate rate of carbon dioxide assimilation to changes in the activity of several enzymes of the system. The model requires measurements of the steady-state rate of carbon dioxide assimilation, the concentrations of several intermediates in the photosynthetic system, and the concentration of the active site of ribulose 1,5-bisphosphate carboxyalse/oxygenase (Rubisco). It is shown that in sunflowers (Helianthus annuus L.) at photon flux densities that are largely saturating for the rate of photosynthesis, the steady-stete rate of carbon dioxide assimilation is most sensitive to Rubisco activity and, to a lesser degree, to the activities of the stromal fructose, 6-bisphosphatase and the enzymes catalysing sucrose synthesis. The activities of sedoheptulose 1,7-bisphosphatase, ribulose 5-phosphate kinase, ATP synthase and the ADP-glucose pyrophosphorylase are calculated to have a negligible effect on the flux under the high-light conditions. The utility of this analysis in developing simpler models of photosynthesis is also discussed.Abbreviations c i intercellular CO2 concentration - C infP supJ control coefficient for enzyme P with respect to flux J - DHAP dihydroxyacetonephosphate - E4P erythrose 4-phosphate - F6P fructose 6-phosphate - FBP fructose 1,6-bisphosphate - FBPase fructose 1,6-bisphosphatase - G3P glyceraldehyde 3-phosphate - G1P glucose 1-phosphate - G6P glucose 6-phosphate - Pi inorganic phosphate - PCR photosynthetic carbon reduction - PGA 3-phosphoglyceric acid - PPFD photosynthetically active photon flux density - R n J response coefficient for effector n with respect to flux J - R5P ribose 5-phosphate - Rubisco ribulose 1,5-bisphosphate carboxylase/oxygenase - Ru5P ribulose 5-phosphate - RuBP ribulose 1,5-bisphosphate - S7P sedoheptulose 7-phosphate - SBP sedoheptulose 1,7-bisphosphate - SBPase sedoheptulose 1,7-bisphosphatase - SPS sucrose-phosphate synthase - Xu5P xylulose 5-phosphate - n P elasticity coefficient for effector n with respect to the catalytic velocity of enzyme P This research was funded by an Australian Research Council grant to I.E.W. and was undertaken during a visity by K.A.M. to the James Cook University of North Queensland. The expert help of Glenys Hanley and Mick Kelly is greatly appreciated.  相似文献   

19.
A model of heat transfer during grinding in vertical multi-disk perl mills has been proposed. Heat transfer intensity in such mills depends on thermal resistance in a boundary layer formed at the inner surface of mill tank wall. The layer thickness changes depending on process variables. Results obtained are presented in the form of a dimensionless correlation equation.List of Symbols C ball filling of the mill, - c pw specific heat of cooling water, kJ/(kg K) - d disk diameter, m - d k ball diameter, m - D inner diameter of the mill tank, m - G w mass flow rate of cooling water, kg/s - h distance between impeller disks, m - n revolutions frequency of the impeller shaft, s–1 - q heat flux density, kW/m2 - Q c total heat energy emitted in the mill, W - T temperature, K - T w1 temperature of cooling water at the cooling jacket inlet, K - T w2 cooling water temperature at the outlet, K - T m average temperature inside the mill, K - T s average temperature of the tank wall, K - u peripheral speed of the impeller disk, m/s - heat transfer coefficient, kW/(m2K) - boundary layer thickness, m - porosity of the lying bed, - m porosity of the suspended bed, - c liquid dynamic viscosity, Pa s - cs liquid dynamic viscosity at wall temperature, Pa s - c thermal conductivity coefficient of liquid, W/(mK) - c liquid density, kg/m3 - s solid density, kg/m3 Dimensionless Numbers Reynolds number for mixing process - Reynolds number for liquid parameters - Nusselt number for liquid parameters - Prandtl number for liquid parameters - modified Euler number  相似文献   

20.
UVM (ultravioletmodulation of mutagenesis) is a recently describedrecA-independent, inducible mutagenic phenomenon in which prior UV irradiation ofEscherichia coli cells strongly enhances mutation fixation at a site-specific 3-N4-ethenocytosine (C) lesion borne on a transfected single-stranded M13 DNA vector. Subsequent studies demonstrated that UVM is also induced by alkylating agents, and is distinct from both the SOS response and the adaptive response to alkylation damage. Because of the increasing significance being attributed to oxidative DNA damage, it is interesting to ask whether this class of DNA damage can also induce UVM. By transfecting M13 vector DNA bearing a site-specificC lesion into cells pretreated with inducing agents, we show here that the oxidative agent H2O2 is a potent inducer of UVM, and that the induction of UVM by H2O2 does not requireoxyR-regulated gene expression. UVM induction by H2O2 appears to be mediated by DNA damage, as indicated by the observation of a concomitant reduction in cellular toxicity and UVM response in OxyRc cells. Available evidence suggests that UVM represents a generalized cellular response to a broad range of chemical and physical genotoxicants, and that DNA damage constitutes the most likely signal for its induction.  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