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.
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   

2.
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   

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 main carotenoid of Flavobacterium strain R1560 has been identified as (3R,3R)-zeaxanthin. Also present were small amounts of 15-cis-phytoene, phytofluene, -carotene (7,8,7,8-tetrahydro-, -carotene plus 7,8,11,12-tetrahydro-, -carotene), neurosporene, lycopene, -zeacarotene, -carotene, -carotene, -cryptoxanthin, rubixanthin, 3-hydroxy--zeacarotene and several apo-carotenals. Zeaxanthin production was inhibited by nicotine (10 mM), and lycopene and rubixanthin accumulated. The biosynthesis of zeaxanthin is discussed in terms of pathways and also of half-molecule reaction sequences. The presence of zeaxanthin may be a characteristic of a group of Flavobacterium species, and may thus be useful in the taxonomic classification of these organisms.  相似文献   

5.
A model of membrane potential-dependent distribution of oxonol VI to estimate the electrical potential difference across Schizosaccharomyces pombe plasma membrane vesicles (PMV) has been developed. was generated by the H+-ATPase reconstituted in the PMV. The model treatment was necessary since the usual calibration of the dye fluorescence changes by diffusion potentials (K+ + valinomycin) failed. The model allows for fitting of fluorescence changes at different vesicle and dye concentrations, yielding in ATP-energized PMV of 80 mV. The described model treatment to estimate may be applicable for other reconstituted membrane systems.  相似文献   

6.
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   

7.
Summary Leaf water potentials, osmotic properties and structural characteristics were examined in the Australian tropical rainforest tree species, Castanospermum australe. These features were compared for individuals growing in the understorey and canopy of the undisturbed forest and in an open pasture from which the forest had been cleared. Leaf water potentials during the day declined to significantly lower values in the open-grown and canopy trees than in the understorey trees. During most of the day the opengrown tree experienced the lowest water potentials. These differences were paralleled by significant differences in tissue osmotic properties. The tissue osmotic potential at full hydration was lowest in the open-grown tree (-1.80 MPa), intermediate in the canopy trees (-1.38 MPa), and highest in the understorey trees (-0.80 MPa). As a result, the degree to which high and positive turgor pressures were maintained as water potentials declined was highest in the open-grown tree, intermediate in the canopy trees, and lowest in the understorey trees. The differences in tissue osmotic properties between individuals in the three crown positions were paralleled, in turn, by differences in leaf structual characteristics. Relative to leaves of the canopy and open-grown trees, leaves of the understorey trees had significantly larger epidermal cells with thinner cell walls, larger specific leaf areas and turgid weight: dry weight ratios, and a higher proportion of intercellular air space.Abbreviations 1 Leaf tissue water potential - min Lowest value of 1 during the day ( noon) - P=0 1 zero turgor - R Relative water content - P Tissue turgor pressure - Tissue osmotic potential - 0 at full hydration  相似文献   

8.
The fluorescence of the voltage sensitive dye, diS-C3-(5), has been analyzed by means of synchronous excitation spectroscopy. Using this rather rare fluorescence technique we have been able to distinguish between the slightly shifted spectra of diS-C3-(5) fluorescence from cells and from the supernatant. It has been found that diS-C3-(5) fluorescence in the supernatant can be selectively monitored at exc = 630 nm and em= 650 nm, while the cell associated fluorescence can be observed at exc= 690 nm and em = 710 nm. A modified theory for the diSC3-(5) fluorescence response to the membrane potential is presented, according to which a linear relationship exists between the logarithmic increment of the dye fluorescence intensity in the supernatant, In I/I°, and the underlying change in the plasma membrane potential, p=pp. The theory has been tested on human myeloid leukemia cells (line ML-1) in which membrane potential changes were induced by valinomycin clamping in various K+ gradients. It has been demonstrated that the membrane potential change, p,can be measured on an absolute scale. Offprint requests to: J. Plasek  相似文献   

9.
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   

10.
Summary Environmental and water relations parameters during fall were monitored for six conifer tree species common to the central Rocky Mountains growing naturally at the same location (Pinus contorta, Pinus ponderosa, Pinus flexilus, Pseudotsuga menziesii, Abies lasiocarpa, Picea engelmannii). Subsequent to what appeared to be the beginning of seasonal stomatal closure, leaf conductance to water vapor declined sharply following the onset of freezing air temperatures at night. A coincident rapid decline in morning xylem pressure potentials (p) also occurred which resulted in values that were considerably below afternoon p. Continuing decreases in maximum leaf conductance during the day were highly correlated with corresponding decreases in minimum nocturnal air temperatures of the preceding night. By mid-December, morning p returned to values very near afternoon p and were only slightly lower than before the onset of subfreezing nights. A preliminary model is proposed which interprets the qualitative interaction between air and soil temperatures, soil and plant water potentials, and leaf conductance during seasonal stomatal closure in fall.  相似文献   

11.
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   

12.
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   

13.
Book reviews     
Consider the perturbed harmonic oscillator Ty=-y+x2y+q(x)y in L2(), where the real potential q belongs to the Hilbert space H={q, xq L2()}. The spectrum of T is an increasing sequence of simple eigenvalues n(q)=1+2n+n, n 0, such that n 0 as n. Let n(x,q) be the corresponding eigenfunctions. Define the norming constants n(q)=limxlog |n (x,q)/n (-x,q)|. We show that for some real Hilbert space and some subspace Furthermore, the mapping :q(q)=({n(q)}0, {n(q)}0) is a real analytic isomorphism between H and is the set of all strictly increasing sequences s={sn}0 such that The proof is based on nonlinear functional analysis combined with sharp asymptotics of spectral data in the high energy limit for complex potentials. We use ideas from the analysis of the inverse problem for the operator -ypy, p L2(0,1), with Dirichlet boundary conditions on the unit interval. There is no literature about the spaces We obtain their basic properties, using their representation as spaces of analytic functions in the disk.  相似文献   

14.
Summary Determinations of current-voltage relationships are widely employed in the characterization of epithelial sodium transport. In order to determine the protocol dependence of transport parameters in the toad urinary bladder, studies were carried out in the presence and absence of amiloride, an inhibitor of active sodium transport. With symmetric positive and negative perturbations of the transepithelial electrical potential difference (0±100 mV) for 30 sec, the amiloride-sensitive current-voltage (i a -) relationship was near linear over the range –75+100 mV, indicating constancy of the conductance a and the apparent electromotive force E Na, lumped parameters of the standard electrical equivalent circuit model of the active transport system. With a reverse protocol (±1000 mV) or 15 min perturbations thei a - relationships were highly nonlinear. Nonlinearity reflected voltage dependence of parameters: perturbations that increased active transport decreased E Na and increased a, as evaluated from 10 sec perturbations of ; slowing of active transport produced the converse changes. These effects are usefully analyzed in both quasi-steady states and true steady states by means of a detailed equivalent circuit incorporating the significant ionic currents across each plasma membrane. Precise understanding of the significance of a and E Na will require characterization of the partial ionic conductances on perturbation of .  相似文献   

15.
Summary In the CAM plant Kalanchoë daigremontiana, kept in an environmental rhythm of 12 h L: 12 h D in a growth chamber at 60% relative humidity and well watered in the root medium, decreasing water potentials and osmotic potentials of the leaves are correlated with malate accumulation in the dark. In the light increasing water and osmotic potentials ( W and S ) are associated with decreasing malate levels. Transpiratory H2O loss is high in dark and low in light.In continuous light, the CAM rhythm rapidly disappears in the form of a highly damped endogenous oscillation. Malate levels, and water and osmotic potentials of the leaves remain correlated as described above. However, transpiration is very high as malate levels decrease and water and osmotic potentials increase.It can concluded, that water relation parameters like total water potential ( W ) and osmotic potential ( S ) change in close correlation with changes of malic acid levels. As an important osmotically active solute in CAM plants, malic acid appears to affect water relations independently of and in addition to transpiration. The question remains open, whether turgor ( P ) is involved in CAM regulation in intact plants in a similar way as it determines malate fluxes in leaf slices.Abbreviations CAM Crassulacean Acid Metabolism - L Light - D Dark  相似文献   

16.
Ornithine decarboxylase (ODC) plays an essential role in various biological functions, including cell proliferation, differentiation and cell death. However, how it prevents the cell apoptotic mechanism is still unclear. Previous studies have demonstrated that decreasing the activity of ODC by difluoromethylornithine (DFMO), an irreversible inhibitor of ODC, causes the accumulation of intracellular reactive oxygen species (ROS) and cell arrest, thus inducing cell death. These findings might indicate how ODC exerts anti-oxidative and anti-apoptotic effects. In our study, tumor necrosis factor alpha (TNF-) induced apoptosis in HL-60 and Jurkat T cells. The kinetic studies revealed that the TNF- -induced apoptotic process included intracellular ROS generation (as early as 1 h after treatment), the activation of caspase 8 (3 h), the cleavage of Bid (3 h) and the disruption of mitochondrial membrane potential ( m) (6 h). Furthermore, ROS scavengers, such as glutathione (GSH) and catalase, maintained m and prevented apoptosis upon treatment. Putrescine and overexpression of ODC had similar effects as ROS scavengers in decreasing intracellular ROS and preventing the disruption of m and apoptosis. Inhibition of ODC by DFMO in HL-60 cells only could increase ROS generation, but did not disrupt m or induce apoptosis. However, DFMO enhanced the accumulation of ROS, disruption of m and apoptosis when cells were treated with TNF- . ODC overexpression avoided the decline of Bcl-2, prevented cytochrome c release from mitochondria and inhibited the activation of caspase 8, 9 and 3. Overexpression of Bcl-2 maintained m and prevented apoptosis, but could not reduce ROS until four hours after TNF- treatment. According to these data, we suggest that TNF- induces apoptosis mainly by a ROS-dependent, mitochondria-mediated pathway. Furthermore, ODC prevents TNF- -induced apoptosis by decreasing intracellular ROS to avoid Bcl-2 decline, maintain m, prevent cytochrome c release and deactivate the caspase cascade pathway.  相似文献   

17.
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   

18.
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   

19.
Quemada  M.  Cabrera  M.L. 《Plant and Soil》1997,189(1):127-137
A better understanding of the effect of temperature (T) and moisture on soil microbial activity should improve our ability to predict N mineralization from soil organic matter and crop residues. The objective of this study was to evaluate the effects of water potential () and T on C and N mineralization from unamended Cecil loamy sand soil (clayey, kaolinitic, thermic Typic Kanhapludult) and from crimson clover (Trifolium incarnatum L.) residues applied on the soil surface. Cecil soil was packed into acrylic plastic cylinders, adjusted to -5.0, -1.5, -0.03, or -0.003 MPa, treated with clover residues on the surface or left unamended, and incubated at 10, 20, 28, or 35°C for 21 d. Headspace gas samples for CO2 and N2O determinations were taken periodically and NH3 evolved was trapped. Inorganic N in soil and residue extracts was analyzed after 21 d. When increased from -5.0 to -0.003 MPa, total CO2 evolved from unamended soil increased linearly with ln(-), whereas total CO2 evolved from clover residue increased exponentially with . In both cases the effect of was enhanced as T increased. Two-dimensional (T, ) equations were developed to describe these effects. Apparent net mineralized N from the clover residue increased with until it reached a maximum between -0.5 and -0.03 Mpa.  相似文献   

20.
A. K. Knapp 《Oecologia》1984,65(1):35-43
Summary The water relations and growth of three tallgrass prairie species Panicum virgatum, Andropogon gerardii and A. scoparius were examined in irrigated and unwatered prairie in eastern Kansas (USA). Measurements of the osmotic potential at full turgor, 100 , at zero turgor, 0, and growth of vegetative and reproductive tillers were made in a year with above-normal precipitation and a drought year to evaluate: 1) the ability of these grasses to osmotically adjust in response to water stress and 2) the effect of drought or supplemental water on growth of these species. Although these grasses adjusted osmotically even in the wet year, the degree of adjustment of 100 and 0 in the drought year was relatively large (0.60–0.78 MPa and 0.88–1.34 MPa, respectively) compared to reports for other species. Seasonal minimum values of 100 and 0 for these grasses in the drought year were lower than in most mesic species and seasonal fluctuations in 100 and 0 were greater than reported for most mesic or xeric species. The relatively frequent occurrence of drought in sub-humid tallgrass prairies may partially explain the greater than expected magnitude of osmotic adjustment in these grasses.Irrigation in the wet year increased reproductive biomass in the mesic grass P. virgatum, but had no effect on A. gerardii or the more xeric grass A. scoparius. However, irrigation in the drought year increased maximum shoot biomass in all three grasses significantly with the largest increase in P. virgatum. Reproduction in P. virgatum was also increased more by irrigation in the drought year compared to the other grasses. Irrigation did not increase season's end production of A. gerardii in the wet year, but in the drought year production was 28% greater in irrigated than unwatered prairie. The combination of these water relations and growth responses of the three grasses to wetter than normal and drought years supports their reported distribution along a moisture gradient in tallgrass prairies.  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