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-stressed maize (Zea mays L.) leaves showed a large decrease in leaf conductance during photosynthesis. Net CO2 uptake and evaporation declined fast at mild stress (=–0.6 to –1.0 MPa) and slower at more severe stress (=–1.0 to -1.2 MPa), whereas the CO2 concentration in the intercellular spaces (Ci) did not drop to the CO2 compensation point. The activities of the enzymes of photosynthetic carbon metabolism tested in this study dropped by approx. 30% at =-1.2 MPa. Glutamine synthetase activity was unaffected by water stress, whereas the activity of nitrate reductase was almost completely inhibited. The decline of enzyme activities in relation to was correlated with a concomitant decrease in the content of total soluble protein of the stressed leaves. The total leaf pools of malate, pyruvate and oxaloacetate decreased almost linearly in relation to , thus obviously contradicting the almost constant Ci. In comparison to the controls (=0.6 MPa) the content of citrate and isocitrate increaed markedly at =-0.9 MPa and decreased again at =-1.2 MPa.Abbreviations PCR photosynthetic carbon reduction cycle - PCO photosynthetic carbon oxidation cycle - PEP phosphoenolypyruvate - RuBP ribulose-1,5-bisphosphate  相似文献   

2.
Summary Leaf water potential ( l ), osmotic potential ( s ), pressure potential ( p , turgor pressure), relative water content (R) and their interrelationships were determined for a xeric grass (Agropyron dasystachyum) found in the grasslands of Canada. Thermocouple psychrometers were used to measure l and s ; p was obtained by subtraction. l dropped from near 0 bars to about-28 bars as R went from 90% to 75%. R greater than 90% was not observed, perhaps because of a systematic error in determination of turgid water content. R remained relatively high in A. dasystachyum, even at low l . The slope of the l -R relationship was similar to other species which are generally considered to be drought tolerant. p as high as 14 bars was observed. Most of the decrease in l was accounted for by a decline in p . The ability of A. dasystachyum to adjust to fluctuating water stress over the growing season is probably as much related to changes in tissue structure and turgor relationships as to simple changes in osmotic potential.  相似文献   

3.
Pseudo-peptide bond inhibitors (-bond inhibitors) and peptide-aldehyde inhibitors of atrial granule serine proteinase, the candidate processing enzyme of pro-atrial natrieuretic factor, are prepared in high yield and purity by novel synthetic routes. The -bond compounds retain essential residues for enzyme binding, but place the enzyme inhibition site in the midst of the peptide sequence. Thus, Bz-APR--LR and Bz-APR--SLRR can be considered readthrough inhibitors of atrial granule serine proteinase. The most potent -peptide, Bz-APR--SLRR (IC50=250 M), is about fivefold less potent than the best peptide-aldehyde inhibitor (EACA-APR-CHO), and both the -bond and peptide-aldehyde compounds are competitive, reversible inhibitors of the enzyme. The -bond peptides containing two C-terminal Arg residues are three-to tenfold more potent than the analogous compounds containing only one C-terminal Arg residue, confirming the importance of both Arg residues in the enzyme processing recognition site. As expected, because of their moderate potencies, the -peptides are not useful affinity ligands for purification of atrial granule serine proteinase, but both peptide aldehydes are effective affinity ligands [Damodaran and Harris (1995),J. Protein Chem., this issue].Abbreviations AGSP atrial granule serine proteinase - ANF atrial natriuretic factor - Bz benzoyl - DIEA diisopropylethylamine - DIPCDI diisopropylcarbodiimide - DMF dimethylformamide - DMSO dimethylsulfoxide - EACA 6(e)-aminocaproic acid - EtOAc ethyl acetate - HEPES N-2-hydroxyethylpiperazine-N-propanesulfonic acid - HOBt N-hydroxybenzotriazole - HPLC high-performance liquid chrornatography - NMR nuclear magnetic resonance - PEG polyethylene glycol-3350 - PyBOP benzotriazole-1-yl-oxy-trispyrrolidino-phosphonium-hexafluorophospate - TEA triethylamine - TFA trifluoroacetic acid - THF tetrahydrofuran - TLC thin-layer chromatography - UV ultraviolet - pseudo-peptide bond -CH2-NH-. Single-letter abbreviations are used to denote amino acids  相似文献   

4.
The synthesis of 2-methyl-5-amino-4-oxo-3-sulfonyl esters, potentialprecursors of Xaa[COCH2]Ala, Xaa[E-CH=CH]Ala andXaa[CH2CH2]Ala pseudodipeptides, has been investigated byalkylation of aminoacid-derived -ketosulfones with ethyl 2-bromo- or2-triflyloxypropionate in different basic conditions. Yields in 2-methyl-5-amino-4-oxo-3-sulfonyl esters are low but starting -ketosulfones are recovered in good yield.  相似文献   

5.
Photosynthetic potential of isolated chloroplasts was investigated during in situ water deficits. An eight day stress cycle imposed on spinach plants reduced leaf w by 0.57MPa, and leaf by 0.50MPa, resulting in partial turgor maintenance during the stress cycle. Pressure/volume curves confirmed the occurrence of osmotic adjustment. Leaf depression was associated with an altered response of chloroplasts to low in vitro. Optimum reaction medium for photosynthesis shifted from –1.04 to –1.57MPa, and low was not as inhibitory to photosynthesis of plastids pre-exposed to stress in situ. These data indicate that chloroplasts acclimate to low external in response to leaf water deficits. This response was still evident four days after a stress cycle ended, but was nearly reversed eight days after stress. Repeated stress cycles in situ did not increase the degree of chloroplast acclimation to low in vitro. Fast dehydration of leaves did not induce this apparent chloroplast acclimation.Abbreviations osmotic potential - w water potential - PEG polyethylene glycol 8000 - MPa megapascals  相似文献   

6.
Summary The potential-sensitive response mechanism of 3,3-dipropylthiodicarbocyanine iodide (diS-C3-(5)) was examined based on our previous model of diS-C3-(5) interaction with brush border membrane vesicles (BBMV) in the absence of a membrane potential. The model contained binding (6 msec), reorientation (30 msec), dimerization (<10 nsec), and translocation (1 sec) reaction steps (Cabrini & Verkman, 1986.J. Membrane Biol. 90:163–175). Transmembrane potentials () were induced in BBMV by K+ gradients and valinomycin. Steady-state diS-C3-(5) fluorescence (excitation 622 nm, emission 670 nm) increased linearly with . The reorientation and translocation reaction steps were resolved by the stopped-flow technique as a biexponential decrease in fluorescence following mixture of diS-C3-(5) with BBMV at varying . The fractional amplitude of the faster exponential increased from 0.36 to 0.73 with increasing (–17 to 87 mV); the time constant for the faster exponential (30–35 msec) was independent of . There were single exponential kinetics (0.5–1.5 sec) for diS-C3-(5) fluorescence response to a rapid (<2 msec) change in in the absence of a diS-C3-(5) concentration gradient. These results, and similar findings in placental brush border vesicles, red cell vesicles, and phosphatidylcholine vesicles, support a translocation mechanism for diS-C3-(5) response, where induced membrane potentials drive diS-C3-(5) redistribution between sites at the inner and outer membrane leaflets, with secondary effects on diS-C3-(5) dimerization and solution/membrane partitioning. Fluorescence lifetime and dynamic depolarization measurements showed no significant change in diS-C3-(5) rotational characteristics or in the polarity of the diS-C3-(5) environment with changes in . Based on the experimental results, a mathematical model is developed to explain the quantitative changes in diS-C3-(5) fluorescence which accompany changes in at arbitrary dye/lipid ratios.  相似文献   

7.
This study employed an intensive sampling regime in which leaf gas exchange and tissue-water relations were measured simultaneously on the same leaf at midday on 19 tree species from three distinct forest communities during wet (1990) and dry (1991) growing seasons. The study sites were located on a xeric barrens, a misic valley floor, and a wet-mesic floodplain in central Pennsylvania, United States. The xeric, mesic, and wetmesic sties had drought-related decreases in gravimetric soil moisture of 53, 34 and 27%, respectively. During the wet year, xeric and mesic communities had high seasonal mean photosynthetic rates (A) and stomatal conductance of water vapor (g wv) and low midday leaf water potential (), whereas the wet-mesic community had low A and g wv and high midday . The mesic and wet-mesic communities had dry year decreases in predawn , g wv and A with the greatest drought effect occurring in the mesic community. Regression analysis indicated that species from each site that exhibited high wet-year A and g wv tended to have low midday . This trend was reversed only in the mesic community in the drought year. Despite differences in midday , all three communities had similar midday leaf turgor pressure (p) in the wet year attributable to lower osmotic potential at zero turgor ( 0 ) with increasing site droughtiness. Lower wet year 0 in the xeric community was due to low symplast volume rather than high solute content. Species with the lowest 0 in the wet year often did not have the lowest 100 possibly related to differences in tissue elasticity. Moreover, increased elasticity during drought may have masked osmotic adjustment in 100 but not in 0 , via dilution of solutes at full hydration in some species. Despite the sampling regime used, there were no relationships between gas exchange and osmotic and elastic parameters that were consistently significant among communities or years. This result questions the universal, direct effect of osmotic and elastic adjustments in the maintenance of photosynthesis during drought. By including a large number of species, this study provided new insight to the ecophysiology of contrasting forest communities, and the community-wide impact of drought on contrasting sites.  相似文献   

8.
The mitochondrial membrane potential (deltapsi(m)) in apoptosis; an update   总被引:14,自引:0,他引:14  
Mitochondrial dysfunction has been shown to participate in the induction of apoptosis and has even been suggested to be central to the apoptotic pathway. Indeed, opening of the mitochondrial permeability transition pore has been demonstrated to induce depolarization of the transmembrane potential (m), release of apoptogenic factors and loss of oxidative phosphorylation. In some apoptotic systems, loss of m may be an early event in the apoptotic process. However, there are emerging data suggesting that, depending on the model of apoptosis, the loss of m may not be an early requirement for apoptosis, but on the contrary may be a consequence of the apoptotic-signaling pathway. Furthermore, to add to these conflicting data, loss of m has been demonstrated to not be required for cytochrome c release, whereas release of apoptosis inducing factor AIF is dependent upon disruption of m early in the apoptotic pathway. Together, the existing literature suggests that depending on the cell system under investigation and the apoptotic stimuli used, dissipation of m may or may not be an early event in the apoptotic pathway. Discrepancies in this area of apoptosis research may be attributed to the fluorochromes used to detect m. Differential degrees of sensitivity of these fluorochromes exist, and there are also important factors that contribute to their ability to accurately discriminate changes in m.  相似文献   

9.
Summary Plant water relations and shoot growth rate of shrubs resprouting after fire or unburnt were measured in a semi-arid poplar box (Eucalyptus populnea) shrub woodland of eastern Australia. In vegetation unburnt for about 60 years, the dawn xylem water potential (x) of the dominant shrub species was about-1.0 MPa when the soil was wet and-8.0 MPa when the soil was very dry. At any one time, the dominant shrub species,Eremophila mitchellii, E. sturtii, Geijera parviflora andCassia nemophila, were similar in x butAcacia aneura andDodonaea viscosa were consistently higher in x than this group when the soil was moist and lower when the soil was dry. The dominant tree species,Eucalyptus populnea andE. intertexta, appeared to have access to additional water beneath the hardpan which is located 60–80 cm below the surface. When shrubs were under extreme water stress (x of-8 MPa), the trees had a x of-3 to-3.6 MPa. Following a fire, both x and leaf stomatal conductance (g s) of resprouting shrubs were higher for about 5 years than comparable-aged unburnt vegetation, with relative differences in x increasing with drought stress. Elongation rate of resprouts was positively linked to prefire shrub height in 3 of 4 species. However, shrubs resprouting after high intensity fires had substantially higher rates of shoot elongation than after low intensity fires which were in turn higher than for foliar expansion of unburnt shrubs. It is concluded that the growth rate of resprouting shrubs is primarily determined by physiological/ morphological factors associated with plant size but is also assisted by greater availability of water and possibly nutrients for a period after fire.  相似文献   

10.
The inside-out vesicles of plasma membranes were isolated from pumpkin stem cells, and the kinetics of sucrose efflux induced by the K+-diffusion potential (D) was studied by measuring light transmission. Two phases differing in their rates and duration were identified in D-dependent changes of light transmission. The increase in Delevated the rate and magnitude of the fast phase in light transmission changes but did not markedly affect the rate of the slow phase. These two phases also differed in their sensitivity to inhibitors and to changes in sucrose concentration in the external medium. Measurements of Dduring sucrose transport by means of the fluorescence probe dis-C3-(5) revealed differences in the magnitude of Dand its stability in vesicles loaded with sucrose and mannitol, as well as under the action of inhibitors. The two-phase dependence of sucrose efflux from vesicles on the applied diffusion potential is discussed in the context of modern concepts on the functioning of sucrose carriers in the membranes.  相似文献   

11.
R. J. Fellows  J. S. Boyer 《Planta》1976,132(3):229-239
Summary Changes in membrane integrity, conformation and configuration, and in photosystem II (PS II) activity (measured as dichloroindophenol photoreduction) of sunflower (Helianthus annuus L.) chloroplasts were studied after leaf tissue had been desiccated to various water potentials ( w ). Fixatives for electron microscopy were adjusted osmotically to within 1 bar of the w of the tissue to prevent rehydration during fixation. PS II activity decreased to 50% of the control activity at a w of-26 bar. At this w , leaf viability was being lost but there was virtually no loss of integrity of the thylakoid lamellar system. Even at extreme w (below-100 bar), thylakoids retained much structural detail but were less stained. At-26 bar, intrathylakoid spacing (configuration) and lamellar thickness (conformation) were decreased in vivo. Upon isolation of the plastids, the differences in configuration disappeared but the differences in conformation remained. The decreases in membrane conformation and PS II activity both, in vivo and in vitro suggest that alterations in conformation may cause decreases in chloroplast activity at w as low as-26 bar. Since structural detail was maintained, however, previous observations of altered membrane integrity, which involved tissue fixed without osmotic support, may have been affected by tissue rehydration during fixation.Abbreviations DCIP sodium 2,6-dichloroindophenol - PS II photosystem II - w leaf water potential  相似文献   

12.
Summary Electrical potential differences across the plasma membrane () of the yeastPichia humboldtii were measured with microelectrodes (filled with 0.1m KCl) inserted into cells immobilized in microfunnels. The registered signals were reproducible and stable for several minutes. On attainment of stable reading for the specific membrane resistanceR sp was determined by applying square-current pulses to the preparation. Both andR sp were pH dependent and displayed equal but opposite deflection, reaching its maximal value of –88±9 mV (n=13) andR sp its minimal value of 10 k·cm2 (maximal conductance) at pH 6.5. Uncouplers and the polyene antibiotic nystatin depolarized the cells, decreasing to –21±15 mV (n=10) with concomitant decrease ofR sp. Comparison of values from microelectrode measurements with those calculated from the steady-state distribution of tetraphenylphosphonium ions agreed within 10 mV under all physiological conditions tested, except at pH values above 7.0. During microelectrode insertion transient voltage signals (a few msec long) were detected by means of an oscilloscope. These voltage signals were superimposed on the stable recordings described above. These short voltage signals disappeared in uncoupled cells. The closely related values obtained by two independent methods (direct measurements with microelectrodes and calculation from steady-state distribution of a lipophilic cation) provide evidence that these values reffect the true membrane potential of intact cells.  相似文献   

13.
Summary The total carbon 13C values of two C3 halophytes,Salicornia europaea L. ssp.rubra (Nels.) Breitung andPuccinellia muttalliana (Schultes) Hitch., native to inland saline areas of Alberta, Canada, were determined for plants grown under controlled conditions of supplied NaCl in the nutrient solution, and for plants found growing in the field. Field specimens were collected along line transects which ran from areas of high salinity to areas of low salinity across the pattern of species zonation. The 13C value of the two species seemed to reflect the water potential of the soil ( w soil ) as measured arbitrarily at a depth of 10 cm, becoming less negative as the w soil decreased. Over a linear distance of 5.55 m,S. europaea spp.rubra showed a shift of +5.3 as the w soil went from-25x102 kPa to a minimum of-73x102 kPa. ForP. nuttalliana, the 13C values differed by 3.4 over a distance of 7.45 m where the maximum difference in w soil was 12.7x102 kPa. However, 13C values ofP. nuttalliana only roughly reflected the spatial trends in w soil at the time of collection. In the growth chamber, the 13C value ofS. europaea ssp.rubra changed by a maximum of +8.0 when the solute potential of the nutrient solution ( w soil ) was dropped from-0.25x102 kPa to-64.25x102 kPa; while the 13C value ofP. nuttalliana changed by a maximum of +10.8 when the w soil was dropped from-0.25x102 kPa to-40.25x102 kPa. Linear regression analyses indicated that the 13C values of both species were strongly correlated (P<0.2%) with w soil . The observed shifts in 12C may represent changes in the mode of photosynthetic CO2 fixation. However, a number of other explanations, some of which are discussed in the text, are also possible. A proper ecophysiological interpretation of such shifts in 13C values of C3 plants awaits a better understanding of the isotope fractionation mechanisms involved.  相似文献   

14.
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   

15.
Summary Pinealectomy of house sparrows on 3L:21D (3 h light per 24 h) resulted in a significant increase in the time between the onset of perch-hopping activity and lights on (on) as well as the time between the offset of activity and lights of (off). The daily variance in on and off was also increased following the removal of the pineal gland. On longer light cycles (i.e., 5L:19D; 7L:17D), neither on or off, nor the variance of on or off was different between sham-pinealectomized and pinealectomized sparrows. Upon returning the birds to an ultrashort light cycle, 1L:23D, off, as well as the variance in on and off were again found to be significantly larger in the pinealectomized birds when compared to sham-operated controls. These results indicate that the effects of pinealectomy on the entrained rhythm of locomotor activity are most pronounced when birds are exposed to a weak entraining agent, such as an ultrashort light: dark cycle. In view of the observation that pinealectomy can alter the phase relationship between activity onset and offset, it is suggested that the pineal gland may be involved in the coupling of the oscillators that regulate activity onset and offset.  相似文献   

16.
Summary M1 is a virulent bacteriophage of Methanobacterium thermoautotrophicum strain Marburg. Restriction enzyme analysis of the linear, 30.4 kb phage DNA led to a circular map of the 27.1 kb M1 genome. M1 is thus circularly permuted and exhibits terminal redundancy of approximately 3 kb. Packaging of M1 DNA from a concatemeric precursor initiates at the pac site which was identified at coordinate 4.6 kb on the circular genome map. It proceeds clockwise for at least five packaging rounds. Headful packaging was also shown for M2, a phage variant with a 0.7 kb deletion at coordinate 23.25 on the map.  相似文献   

17.
Summary Phloretin and other neutral phloretin-like molecules are able to decrease the electrostatic potential within neutral lipid bilayers and monolayers. The relationship between the change in the dipole potential and the aqueous concentration of the molecule is well described by a Langmuir isotherm. From the Langmuir isotherm, the apparent dissociation constants (K D A ) and the maximum dipole potential change ( max) are obtained for the different phloretin-like molecules tested. Considering the phloretin analogs as derivatives of acetophenone containing two kinds of substituents, one on the benzene ring and another on the carbon chain, it is found that (a)K D A is related to the hydrophobicity of the compound and is also a function of the position of the hydroxyl substituent in the ring; (b) from the dependence ofK D A on the length of the acyl chain, it is estimated that the free-energy change is 650 cal/mole CH2; (c) max is not a simple function of the dipole moment of the molecule but depends on the substituent on the carbon chain and on the position and number of hydroxyl groups on the benzene ring; (d) phloretin adsorption parameters are a function of membrane lipid composition. The results are discussed in terms of the effect of these compounds on chloride transport in red blood cells.  相似文献   

18.
Summary Shoot water relations, summer gas exchange response and morphological development of western hemlock [Tsuga heterophylla (Raf.) Sarg.] and western red cedar (Thuja plicata Donn) seedlings were monitored over the first growing season on a coastal reforestation site in British Columbia. In March, osmotic potential (s) at saturation [s(sat)] was –1.98 MPa and turgor loss point [s(tlp)] –2.38 MPa for western hemlock, while western red cedar had –1.45 MPa s(sat) and –1.93 MPa s(tlp). Seasonally s increased through June and then decreased through September, with western hemlock –0.15 to –0.50 MPa lower than western red cedar. Maximum bulk modulus of elasticity (max) for western hemlock was 29.3 MPa in March, decreased to 15.0 MPa in June and increased to 25.0 MPa from July through September, while western red cedar max was 10.6 MPa in March and around 8.0 MPa thereafter. Utilized turgor (T util) for western hemlock was <40% from March through May, 69 to 78% from June through August and 96% in September, while western red cedar T util was 68 to 73% during March and April, 84 to 96% from May through August and 100% in September. Maximum CO2 assimilation rate (A) of western red cedar was more than double western hemlock, and for both species A declined in a linear fashion with increasing vapour pressure deficit (D). Maximum foliage conductance (g wv) declined in a concave manner as D increased in both species, with western red cedar values 50 to 67% greater than western hemlock. Maximum daily g wv declined in a concave manner as predawn shoot water potential (pd) decreased, with maximum daily g wv 1.8 to 3.6 times greater in western red cedar than western hemlock, when pd was –0.25 and –1.4 MPa, respectively. Western red cedar, compared to western hemlock, had a greater increase in A as g wv increased. Eight months after planting, western red cedar seedlings had twice the root growth, measured as root dry weight and root number, of western hemlock.  相似文献   

19.
The development of CAM-type photosynthesis is one of the adaptation mechanisms to severe water deficit. It provides plants with carbon dioxide and permits efficient water spending under extreme environments. In common ice plants, a complete switch from C3 to CAM photosynthesis was observed on the seventh day of salinity (0.5 M NaCl). The indices characterizing this switch were: (1) induction of phosphoenolpyruvate carboxylase; (2) diurnal changes in the organic acid content, which are characteristic of CAM plants, and (3) suppression of transpiration during the daytime. A decrease in the osmotic potential () of the leaf sap, which occurred on the second day of salinity, preceded these changes. After long-term salinity stress (four–five weeks), attained extremely low values (–4.67 MPa), which made possible the water uptake by the root system. The restoration of the balance between cell compartments resulted from the accumulation of compatible solutes in the cytoplasm, proline primarily, which possesses osmoregulatory and stress-protective properties. This means that a complex of adaptive mechanisms is required for the realization of the common ice developmental program under salinity. These mechanisms maintained plant capacity to uptake water and permitted its efficient utilization. They triggered the development of stress-induced CAM-type photosynthesis, maintained the low osmotic potential in the cell sap, regulated the composition of macromolecules in the cell microenvironment, provided for water storage in tissues, and reduced the time of plant development. A comparison between the time-courses of CAM development and a decrease in the transpiration rate permitted us to suggest that a combination of low and CO2 in the leaf cells could serve as a signal for the induction of CAM-dependent gene expression in terrestrial plants.  相似文献   

20.
S. T. C. Wright 《Planta》1977,134(2):183-189
The amount of diffusible ethylene from excised wheat leaves (Triticum aestivum L. cv. Eclipse) increased when they were subjected to water stress. The quantity of ethylene produced was related to the severity of the stress, reaching a maximum at a leaf water potential leaf of approximately-12 bars. Irrespective of the severity of the stress, the maximum rate of ethylene production usually occurred between 135–270 min after applying the stress and then the rate declined. Part of the decline may have been due to an oxygen deficiency in the leaf chambers. In excised water-stressed leaves there was a sigmoid relationship between increasing ethylene and abscisic acid (ABA) levels and decreasing leaf water potential values. The two curves were displaced from each other by approximately 1 bar, with ethylene evolution leading that of ABA accumulation. The maximum rate of increase in ethylene occurred between-8 and-9 bars and for ABA between-9 and-10 bars. A significant increase in the levels of these two plant growth regulators was found when the leaf decreased outside the normal diurnal leaf range by 1 bar for ethylene and 2 bars for ABA. Because of the sigmoid nature of the curves there was no distinct threshold leaf value triggering-off an increase in ethylene or ABA, but with ABA the curve became very steep at a leaf value of-9.3 bars and this could be looked upon as a kind of threshold value.It seems unlikely that the stress-induced ethylene evolution in excised wheat leaves stimulated the accumulation of ABA, because when the leaves were subjected to a substantial water stress (e.g. leaf bars) ABA increased immediately and at a faster rate than ethylene.Abbreviations ABA abscisic acid - GLC gas-liquid chromatography - RWC relative water content - TLC thin-layer chromatography - leaf leaf water potential  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