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 234 毫秒
1.
Bound Water in Durum Wheat under Drought Stress   总被引:1,自引:0,他引:1       下载免费PDF全文
To study drought stress effects on bound water, adsorption isotherms and pressure-volume curves were constructed for two durum wheat (Triticum durum Desf.) cultivars: Capeiti 8 (drought tolerant) and Creso (drought sensitive). Plants were grown under well-watered and water-stressed conditions in a controlled environment. Differential enthalpy (ΔH) was calculated through van't Hoff analysis of adsorption isotherms at 5 and 20°C, which allowed us to determine the strength of water binding. ΔH reached the most negative values at approximately 0.06 gram H2O/gram dry weight and then increased rapidly for well-watered plants (until 0.10 gram H2O/gram dry weight) or more slowly for drought-stressed plants (until 0.15-0.20 gram H2O/gram dry weight). Bound water values from pressure-volume curves were greater for water-stressed (0.17 gram H2O/gram dry weight) than for well-watered plants (0.09 gram H2O/gram dry weight). They may be estimates of leaf moisture content where ΔH reaches the less negative values and hence some free water appears. With respect to the well-watered plants, tightly bound water tended to be less bound during drought, and more free water was observed in cv Creso compared to cv Capeiti 8 at moisture contents >0.10 gram H2O/gram dry weight.  相似文献   

2.
Vertucci CW 《Plant physiology》1989,90(3):1121-1128
In an attempt to correlate freezable water with freezing injury, the thermal behavior of pea (Pisum sativum L.) and soybean (Glycine max L. Merr) seed parts at different moisture contents were compared with survival of the seeds when exposed to low temperatures. Thermal transitions between −150 and 10°C were studied using differential scanning calorimetry. In pea, reduction of germinability, after exposure of seeds to temperatures between − 18 and − 180°C, occurred at a constant moisture content (about 0.33 gram H2O/gram dry weight) regardless of the temperature; this moisture level was above that at which freezable water was first detectable by differential scanning calorimetry (0.26 gram H2O/gram dry weight). In contrast, damage to soybean seeds was observed at progressively lower moisture contents (from 0.33 to 0.20 gram H2O/gram dry weight) when the temperature was decreased from −18°C to −50°C. At −18 and −30°C, moisture contents at which damage to soybean seeds was evident were above that at which freezable water was first detectable (0.23 gram H2O/gram dry weight). However, at −50, −80, and −180°C, damage was evident even when freezable water was not detectable. The data suggest that, while the quantity of water is important in the expression of freezing injury, the presence of freezable water does not account for the damage.  相似文献   

3.
Changes in water content and dry weight were determined in control cells and those induced to cold harden in response to abscisic acid (ABA) treatment (7.5 × 10−5 molar). Bromegrass (Bromus inermis Leyss cv Manchar) cells grown in suspension culture at room temperature (23°C) for 7 days acclimated to −28°C (LT50) when treated with ABA, or to −5°C when untreated. ABA significantly reduced cell growth rates at 5 and 7 days after treatment. Growth reduction was due to a decrease in cell number rather than cell size. When the cell water content was expressed as percent water (percent H2O) or as grams water per gram dry weight (gram H2O/gram dry weight [g DW]), the water content of hardy, ABA-treated cells decreased from 85% to 77% or from 6.4 to 3.3 g H2O/g DW in 7 days. Control cell water content remained static at approximately 87% and 7.5 g H2O/g DW. However, cell water content, expressed as milligrams water per million cells (milligram H2O/106 cells), did not differ in ABA-treated or control cells. The dry matter content of ABA-treated cells, expressed as milligram DW/106 cells increased to 3.3 milligram/106 cells in 7 days, whereas the dry weight of the control cells remained between 1.4 to 2.1 milligrams/106 cells. The osmotic potential of ABA-treated cells decreased by the fifth day while that of control cells increased significantly and then decreased by day 7. Elevated osmotic potentials were not associated with increased ion uptake. In contrast to much published literature, these results suggest that cell water content does not decrease in ABA-treated cells during the induction of freezing tolerance, rather the dry matter mass per cell increased. Cell water content may be more accurately expressed as a function of cell number when accompanying changes to dry cell matter occur.  相似文献   

4.
Vertucci CW 《Plant physiology》1989,90(4):1478-1485
The effect of cooling rate on seeds was studied by hydrating pea (Pisum sativum), soybean (Glycine max), and sunflower (Helianthus annuus) seeds to different levels and then cooling them to − 190°C at rates ranging from 1°C/minute to 700°C/minute. When seeds were moist enough to have freezable water (> 0.25 gram H2O/gram dry weight), rapid cooling rates were optimal for maintaining seed vigor. If the seeds were cooled while at intermediate moisture levels (0.12 to 0.20 gram H2O per gram dry weight), there appeared to be no effect of cooling rate on seedling vigor. When seeds were very dry (< 0.08 gram H2O per gram dry weight), cooling rate had no effect on pea, but rapid cooling rates had a marked detrimental effect on soybean and sunflower germination. Glass transitions, detected by differential scanning calorimetry, were observed at all moisture contents in sunflower and soybean cotyledons that were cooled rapidly. In pea, glasses were detectable when cotyledons with high moisture levels were cooled rapidly. The nature of the glasses changed with moisture content. It is suggested that, at high moisture contents, glasses were formed in the aqueous phase, as well as the lipid phase if tissues had high oil contents, and this had beneficial effects on the survival of seeds at low temperatures. At low moisture contents, glasses were observed to form in the lipid phase, and this was associated with detrimental effects on seed viability.  相似文献   

5.
The dormant cysts of Artemia undergo cycles of hydration-dehydration without losing viability. Therefore, Artemia cysts serve as an excellent intact cellular system for studying the dynamics of water-protein interactions as a function of hydration. Deuterium spin-lattice (T1) and spin-spin (T2) relaxation times of water in cysts hydrated with D2O have been measured for hydrations between 1.5 and 0.1 g of D2O per gram of dry solids. When the relaxation rates (I/T1, I/T2) of 2H and 17O are plotted as a function of the reciprocal of hydration (1/H), an abrupt change in slope is observed near 0.6 g of D2O (or H2 17O)/gram of dry solids, the hydration at which conventional metabolism is activated in this system. The results have been discussed in terms of the two-site and multisite exchange models for the water-protein interaction as well as protein dynamics models. The 2H and 17O relaxation rates as a function of hydration show striking similarities to those observed for anisotropic motion of water molecules in protein crystals.

It is suggested here that although the simple two-site exchange model or n-site exchange model could be used to explain our data at high hydration levels, such models are not adequate at low hydration levels (<0.6 g H2O/g) where several complex interactions between water and proteins play a predominant role in the relaxation of water nuclei. We further suggest that the abrupt change in the slope of I/T1 as a function of hydration in the vicinity of 0.6 g H2O/g is due to a change in water-protein interactions resulting from a variation in the dynamics of protein motion.

  相似文献   

6.
A method for calculating the contraction strain (or the converse stress) to protoplasts of frozen multicellular plant tissues is described. The method requires (i) a nuclear magnetic resonance (NMR) measure of the quantity of bound water per gram dry weight (K), (ii) a gravimetric measure of grams of H2O per gram dry weight (L0), (iii) a measure of solute concentration in non-frozen cells (C0), (iv) an estimate of the specific volume of tissue dry matter (0.625 ml/g), (v) an NMR measure of the fraction of tissue water that is intracellular osmotic water (Pa), and (vi) a measure of the fraction of the dry weight that is cell wall (fcw). This method is a refinement of previous methods that calculate cell contraction strain from four (L0, K, C0, and 0.625) of the above six measurements. Comparison of the calculated protoplast strain to the calculated cell strain indicates that the two measures are quite similar, however, the measure of protoplast strain is, in theory, a more appropriate measure of the freezing strain. It is also demonstrated that derivation of a measure of strain from the parameters controlling it is useful, because it allows one to evaluate the relative contribution of each parameter in preventing the development of strain.  相似文献   

7.
Three tomatoes, Lycopersicon esculentum Mill. cv UC82B, a droughttolerant wild related species, Lycopersicon pennellii (Cor.) D'Arcy, and their F1 hybrid, were grown in containers maintained at three levels of soil moisture. Season-long water use was obtained by summing over the season daily weight losses of each container corrected for soil evaporation. Plant biomass was determined by harvesting and weighing entire dried plants. Season-long water use efficiency (gram dry weight/kilogram H2O) was calculated by dividing the dry biomass by the season-long water use. The season-long water use efficiency was greatest in the wild parent, poorest in the domestic parent, and intermediate (but closer to the wild parent) in the F1 hybrid. Instantaneous water-use efficiency (micromole CO2/millimole H2O) determined by gas exchange measurements on individual leaves was poorly correlated with season-long water use efficiency. However, the relative abundance of stable carbon isotopes of leaf tissue samples was strongly correlated with the season-long water use efficiency. Also, the isotopic composition and the season-long water use efficiency of each genotype alone were strongly negatively correlated with plant dry weight when the dry weight varied as a function of soil moisture.  相似文献   

8.
Laser Raman spectroscopy has been used to investigate the state of water in intact single muscle fibers of the giant barnacle (Balanus nulilus). The spectra in the region of the O-H (or O-2H) stretching modes of water in unfrozen fibers show that there is no appreciable difference between the shape and relative intensity of the Raman bands due to the water molecules located inside a muscle fiber and those of the corresponding bands in the spectrum of pure water. The presence of significant amounts of “structured” intracellular water, greater than approx. 5% of the total water content, in these fibers is thus excluded. The Raman spectra of frozen fibers have also been recorded in order to evaluate the amount of intracellular water which remains unfrozen at temperatures below the normal freezing point of water. We have been able to reproduce these spectra by assuming that the spectrum of a frozen fiber is the sum of the individual spectra of water and ice. To calculate the amount of unfrozen water from these curve fittings, it was also necessary to determine the intensities of the water and ice Raman bands relative to one another. We have found the I(ice)/I(water) ratio is 1.07 ± 0.01 for H2O and 1.05 ± 0.03 for 2H2O With these figures, we have calculated that for a fiber with a normal water content of 80%, 20% of the water molecules remain in the supercooled state at ?5°C, which corresponds to 1 g of water per of fiber dry weight. This amount of bound water was also found to be independent of the water content of the fibers.  相似文献   

9.
Onymacris unguicularis, a fog-basking tenebrionid beetle of the Namib Desert, has mean water influx rates of 49.9 mg H2O/g.d and mean efflux rates of 41.3 mg H2O/g.d with mean mass gain being 10.7 mg/g.d. If only steady-state conditions are considered (no mass change), and passive vapour input subtracted, drinking accounts for 50% of water input. Active beetles must drink in order to maintain water balance, while inactive beetles can maintain water balance either eating seeds or by simply metabolizing fat. Little change is observed in ratios of haemolymph and total body water to dry mass when fogs occur, while significant changes in haemolymph osmotic pressure are associated with fog occurrence.For short periods, O. unguicularis can tolerate dehydration with only slight changes in the ratio of total body water and haemolymph to dry body mass and to haemolymph osmotic pressure. For longer periods however, active beetles must have access to fog water for water balance maintenance. This is probably necessary for reproduction.  相似文献   

10.
We present a method for direct non-optical quantification of dry mass, dry density and water mass of single living cells in suspension. Dry mass and dry density are obtained simultaneously by measuring a cell’s buoyant mass sequentially in an H2O-based fluid and a D2O-based fluid. Rapid exchange of intracellular H2O for D2O renders the cell’s water content neutrally buoyant in both measurements, and thus the paired measurements yield the mass and density of the cell’s dry material alone. Utilizing this same property of rapid water exchange, we also demonstrate the quantification of intracellular water mass. In a population of E. coli, we paired these measurements to estimate the percent dry weight by mass and volume. We then focused on cellular dry density – the average density of all cellular biomolecules, weighted by their relative abundances. Given that densities vary across biomolecule types (RNA, DNA, protein), we investigated whether we could detect changes in biomolecular composition in bacteria, fungi, and mammalian cells. In E. coli, and S. cerevisiae, dry density increases from stationary to exponential phase, consistent with previously known increases in the RNA/protein ratio from up-regulated ribosome production. For mammalian cells, changes in growth conditions cause substantial shifts in dry density, suggesting concurrent changes in the protein, nucleic acid and lipid content of the cell.  相似文献   

11.
Triacylglycerol is used for the production of commodities including food oils and biodiesel fuel. Microalgae can accumulate triacylglycerol under adverse environmental conditions such as nitrogen-starvation. This study explored the possibility of air-drying of green algal cells as a novel and simple protocol for enhancement of their triacylglycerol content. Chlorella kessleri cells were fixed on the surface of a glass fibre filter and then subjected to air-drying with light illumination. The dry cell weight, on a filter, increased by 2.7-fold in 96 h, the corresponding chlorophyll content ranging from 1.0 to 1.3-fold the initial one. Concomitantly, the triacylglycerol content remarkably increased to 70.3 mole% of fatty acids and 15.9% (w/w), relative to total fatty acids and dry cell weight, respectively, like in cells starved of nitrogen. Reduction of the stress of air-drying by placing the glass filter on a filter paper soaked in H2O lowered the fatty acid content of triacylglycerol to 26.4 mole% as to total fatty acids. Moreover, replacement of the H2O with culture medium further decreased the fatty acid content of triacylglycerol to 12.2 mole%. It thus seemed that severe dehydration is required for full induction of triacylglycerol synthesis, and that nutritional depletion as well as dehydration are crucial environmental factors. Meanwhile, air-drying of Chlamydomonas reinhardtii cells increased the triacylglycerol content to only 37.9 mole% of fatty acids and 4.8% (w/w), relative to total fatty acids and dry cell weight, respectively, and a marked decrease in the chlorophyll content, on a filter, of 33%. Air-drying thus has an impact on triacylglycerol synthesis in C. reinhardtii also, however, the effect is considerably limited, owing probably to instability of the photosynthetic machinery. This air-drying protocol could be useful for the development of a system for industrial production of triacylglycerol with appropriate selection of the algal species.  相似文献   

12.
Differential scanning calorimetry was used to study the relationships among drying rate, desiccation sensitivity, and the properties of water in homeohydrous (recalcitrant) seeds of Landolphia kirkii. Slow drying of intact seeds to axis moisture contents of approximately 0.9 to 0.7 gram/gram caused lethal damage, whereas very rapid (flash) drying of excised embryonic axes permitted removal of water to approximately 0.3 gram/gram. The amount of nonfreezable water in embryonic axes (0.28 gram H2O/gram dry mass) did not change with drying rate and was similar to that of desiccation-tolerant seeds. These results suggest that the amount of nonfreezable water per se is not an important factor in desiccation sensitivity. However, flash drying that removed all freezable water damaged embryonic axes. Differences between desiccation-sensitive and -tolerant seeds occur at two levels: (a) tolerant seeds naturally lose freezable water, and sensitive seeds can lose this water without obvious damage only if it is removed very rapidly; (b) tolerant seeds can withstand the loss of a substantial proportion of nonfreezable water, whereas sensitive seeds are damaged if nonfreezable water is removed.  相似文献   

13.
Recent ab initio studies reported in the literature have challenged the mechanistic assignments made on the basis of volume of activation data [1,2]. In addition to that ab initio molecular orbital calculations on hydrated zinc(II)-ions were used to elucidate the general role of this ion in metalloproteins [3]. Due to our interest in both inorganic reaction mechanisms and enzymatic catalysis we started a systematic investigation of solvent exchange processes on divalent zinc-ion using density functional calculations. Our investigations cover aqua complexes of the general form [Zn(H2O)n]2+·mH20 with n=3-6 and m=0-2, where n and m represent the number of water molecules in the coordination and solvation sphere, respectively.The complexes [Zn(H2O)5]2+·2H2O and [Zn(H2O)4]2+·2H2O turnend out to be the most stable zinc complexes with seven and six water molecules, respectively. This implies that a heptacoordinated zinc(II) complex, where all water molecules are located in the co-ordination sphere, should be energetically highly unfavorable and that [Zn(H2O)6]2+ can quite readily push two coordinated water molecules into the solvation sphere. For the pentaqua complex [Zn(H2O)5]2+ only one water molecule is easily lost to the solvation sphere, which makes the [Zn(H2O)4]2+·H2O complex the most favorable in order to consider the limiting dissociative and associative water exchange process of hexacoordinated zinc(II). The dehydration and hydration energies using the most stable zinc(II) complexes [Zn(H2O)4]2+·2H2O, [Zn(H2O)5]2+·2H2O and [Zn(H2O)4]2+·H2O were calculated to be 24.1 and -21.0 kcal/mol, respectively.  相似文献   

14.
《Annals of botany》1997,79(3):291-297
The relationship between the glassy state in seeds and storage stability was examined, using the glass transition curve and a seed viability database from previous experiments. Storage data for seeds at various water contents were studied by Williams–Landel–Ferry (WLF) kinetics, whereas the glass transition curves of seeds with different storage stability were analysed by the Gordon–Taylor equation in terms of the plasticization effect of water on seed storage stability. It was found that the critical temperatures (Tc) for long-term storage of three orthodox seeds were near or below their glass transition temperatures (Tg), indicating the requirement for the presence of the glassy state for long-term seed storage. The rate of seed viability loss was a function of T-Tgat T>Tg, which fitted the WLF equation well, suggesting that storage stability was associated with the glass transition, and that the effect of water content on seed storage was correlated with the plasticization effect of water on intracellular glasses. A preliminary examination suggested a possible link between the glass transition curve and seed storage stability. According to the determined WLF constants, intracellular glasses in seeds fell into the second class of amorphous systems as defined by Slade and Levine (Critical Reviews in Food Science and Nutrition30: 115–360, 1991). These results support the interpretation that the glassy state plays an important role in storage stability and should be a major consideration in optimizing storage conditions.  相似文献   

15.
β-Cyclodextrin (β-CD; cyclomaltoheptaose; cyclohepta-amylose; C42H70O35) crystallises from aqueous solutions of HI and of MeOH in the form of stout prisms, which are isomorphous to each other with monoclinic space-group P21; cell constants for C42H70O35 · 2HI · 8 H2O: a = 21.25(3), b = 10.28(2), c = 15.30(2) Å, β = 113.25(9)°, and Z = 2; and for C42H70O35 · MeOH · 6.5 H2O: a = 21.03(3), b = 10.11(2), c = 15.33(2) Å, β = 111.02(9)°, and Z = 2. X-ray counter data were used to determine the structures of both crystals, which belong to the cage type, with β-CD molecules in nearly identical, “round” shapes. In the HI complex, one I- is located inside, and one outside, the β-CD cavity; in the MeOH complex, the MeOH is within the cavity. The cavity is closed at the O-2,O-3 side by adjacent β-CD molecules, and at the O-6 side by water molecules hydrogen-bonded to the guest and to surrounding β-CD molecules. Interstices between β-CD molecules are filled by water of hydration molecules in distorted co-ordination.  相似文献   

16.
Neutron powder diffraction measurements of fully deuterated protein C-phycocyanin have been made at three temperatures, 295, 200, and 77 K, using dry and partially hydrated samples. The average coherent structure factors and the corresponding radial distribution functions d(r) are determined. The changes in d(r) functions observed in hydrated samples depend strongly on the level of hydration and most of these changes are due to water-protein interactions. At 0.365 gram D2O per gram of protein, the water crystallized into hexagonal ice at 200 K and below, but at 0.175 gram D2O per gram of protein, no crystallization of water was observed. At the higher hydration a peak appears in the radial distribution function which indicates that the average distance of the water molecule in the first hydration shell from the amino acid residues is 3.5 Å.  相似文献   

17.
Differently bound water molecules of dipalmitoylphosphatidylcholine (DPPC)-H2O system were investigated with differential scanning calorimetry (DSC). According to a method previously reported by us, the ice-melting DSC curves of the DPPC-H2O samples of varying water contents were deconvoluted into multiple components, and the ice-melting enthalpies for the individual deconvoluted components were used to estimate average molar ice-melting enthalpies for freezable interlamellar and bulk waters, respectively. With these average molar ice-melting enthalpies, the numbers of differently bound water molecules of the DPPC-H2O system were calculated at varying water contents and were used to construct a water distribution diagram of this system. Furthermore, to evaluate the reliability of the present DSC deconvolution method, 2H-NMR T1 measurements of DPPC-2H2O system were carried out at 5 °C of the gel phase temperature, and components and fractions for differently bound water (2H2O) molecules were estimated from the analysis of nonexponential magnetization recovery curves.  相似文献   

18.
In situ Fourier transform infrared spectroscopy (FTIR) was used in order to obtain more insights in the underlying protective mechanisms upon freezing and drying of ABA-treated tissues of the moss Physcomitrella patens. The effects of different treatments on the membrane phase behaviour, glassy state, and overall protein secondary structure were studied. We found that growth on ABA resulted in the accumulation of sucrose: up to 22% of the tissue on a dry weight basis, compared to only 3.7% in non-ABA-treated tissues. Sucrose functions as a protectant during freezing and drying, but accumulation of sucrose alone is not sufficient for survival. ABA-treated tissue survives a freeze–thaw cycle down to −80 °C only after addition of an additional cryoprotectant (DMSO). Survival correlates with preservation of membrane phase behaviour. We found that ABA-treated P. patens can survive slow but not rapid drying down to water contents as low as 0.02 g H2O per g DW. Rapidly and slowly dried ABA-treated tissues were found to have similar sugar compositions and glass transition temperatures. The average strength of hydrogen bonding in the cytoplasmic glassy matrix, however, was found to be increased upon slow drying. In addition, slowly dried tissues were found to have a higher relative proportion of α-helical structures compared to rapidly dried tissues.  相似文献   

19.
The synthesis of uranium trichloride heptahydrate together with some of its structural, spectroscopic and magnetic properties is reported. The compound possesses the triclinic lattice of LaCl3·7H2O (space group P1). Controlled vacuum thermal dehydration of the substance enabled the preparation of the anhydrous trichloride in gram quantities. Magnetic susceptibilities of polycrystalline samples were measured by the Faraday method in the 6.5–295 K range. The uranium trichloride heptahydrate follows in this region the Curie-Weiss law with C = 1.0839 emu K mol?1 and θ = ?32.7 K.  相似文献   

20.
Summary Cotyledonary somatic embryos ofLarix &#x00D7; leptoeuropaea that developed after various maturation times on media containing abscisic acid showed different frequencies of conversion into plants. Drying of these somatic embryos under high relative humidity (RH) before germination improved plantlet recovery and eliminated differences in the performance of somatic embryos matured for different times. However, dehydration of somatic embryos under 98% RH to a water content below that of zygotic embryos excised from mature seeds (0.97 and 1.36 g H2O/g dry weight, respectively) showed a strong positive correlation between longer maturation time and desiccation tolerance. Drying somatic embryos at 4&#x00B0; C under 59% RH for 1 wk resulted in desiccation to a water content of 0.30 g H2O/g dry weight, which was the closest to the hydration state of zygotic embryos in dried, stored seeds (0.20 g H2O/g dry weight). Under this condition, only somatic embryos matured for 5 wk germinated and produced plantlets at a relatively high frequency (73 and 41%, respectively).  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