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.
Earlier works demonstrated theoretically and experimentally that during gel electrophoresis the mobility μ and the dispersion coefficient Dx [reflecting band broadening; G. W. Slater and J. Noolandi (1985) Physics Review Letters, Vol. 55, pp. 1579–1582; T. A. Duke, A. N. Semenov, and J. L. Viovy (1992) Physics Review Letters, Vol. 69, pp. 3260–3263] depend on the chain length, the electric field, and on the gel concentration. Using a Fluorescence Recovery After Photobleaching setup coupled with an electrophoretic cell, we show that they also depend of the DNA concentrationC. Two regimes are observed. The first is analogous to a “dilute” regime in the gel where μ and Dx are DNA concentration independent. In the second regime, μ and Dx decrease when C increases. These results are explained by DNA-DNA interactions. As expected the C* concentration, under which measurements must be carried out to avoid this effect, is found to be the same as the overlap concentration C* determined in solution. Using concentrations of the studied DNA lower than its C*, μ and Dx show a field dependence in good agreement with the predictions of the Biased Reptation model with Fluctuations. © 1997 John Wiley & Sons, Inc. Biopoly 42: 471–478, 1997  相似文献   

2.
Summary Attenuated total reflection infrared spectroscopy has been used to determine the equilibrium distribution of the peptide antibiotic alamethicinR F30 between dipalmitoyl phosphatidylcholine bilayers and the aqueous environment. The distribution coefficientK=c eq W /c eq M turned out to be concentration dependent, pointing to alamethicin association in the membrane with increasing concentration in the aqueous phase (c eq W ). This concentration was varied within 28 and 310nm, i.e., in a range typical for black film experiments. Furthermore, diffusion coefficients of alamethicin in the hydrophobic phase of the membrane (D M) and across the membrane/water interface (D I) have been estimated from the time course of the equilibration process. It was found that the diffusion rate of the uncharged analogueR F50 is about 10 times higher than that of theR F30 component, exhibiting one negative charge at theC-terminus. The time constants for transmembrane diffusion of alamethicinR F30 varied between 2.2 hr at low concentration and 3.2 hr at higher concentration. The corresponding low concentration value of theR F50 component was found to be 0.25 hr.  相似文献   

3.
Many macromolecular interactions, including protein‐nucleic acid interactions, are accompanied by a substantial negative heat capacity change, the molecular origins of which have generated substantial interest. We have shown previously that temperature‐dependent unstacking of the bases within oligo(dA) upon binding to the Escherichia coli SSB tetramer dominates the binding enthalpy, ΔHobs, and accounts for as much as a half of the observed heat capacity change, ΔCp. However, there is still a substantial ΔCp associated with SSB binding to ssDNA, such as oligo(dT), that does not undergo substantial base stacking. In an attempt to determine the origins of this heat capacity change, we have examined by isothermal titration calorimetry (ITC) the equilibrium binding of dT(pT)34 to SSB over a broad pH range (pH 5.0–10.0) at 0.02 M, 0.2 M NaCl and 1 M NaCl (25°C), and as a function of temperature at pH 8.1. A net protonation of the SSB protein occurs upon dT(pT)34 binding over this entire pH range, with contributions from at least three sets of protonation sites (pKa1 = 5.9–6.6, pKa2 = 8.2–8.4, and pKa3 = 10.2–10.3) and these protonation equilibria contribute substantially to the observed ΔH and ΔCp for the SSB‐dT(pT)34 interaction. The contribution of this coupled protonation (∼ −260 to −320 cal mol−1 K−1) accounts for as much as half of the total ΔCp. The values of the “intrinsic” ΔCp,0 range from −210 ± 33 cal mol−1 °K−1 to −237 ± 36 cal mol−1K−1, independent of [NaCl]. These results indicate that the coupling of a temperature‐dependent protonation equilibria to a macromolecular interaction can result in a large negative ΔCp, and this finding needs to be considered in interpretations of the molecular origins of heat capacity changes associated with ligand‐macromolecular interactions, as well as protein folding. Proteins 2000;Suppl 4:8–22. © 2000 Wiley‐Liss, Inc.  相似文献   

4.
A sedimentation analysis has been used to determine the proportion of protein present as monomer and aggregate in 0.5 and 1.0 g/dl solutions of β-casein A in pH 7 phosphate buffer over the temperature range 10–40°C. The amount and molecular weight of the aggregate increase with temperature; under the conditions used, the aggregation number (n) of β-casein is given approximately by n = 0.6t + 2 with t in degrees centigrade. The concentration of β-casein in monomeric and aggregated states at different temperatures is used to calculate the standard enthalpy of aggregation ΔH° (Van't Hoff) by assuming that β-casein undergoes a cooperative, two-state, micellization process; aggregation is an endothermic process and ΔH° = 66.0 ± 2.6 kJ mol?1. Combination of this ΔH° with the amount of protein calculated to dissociate when 1 g/dl solutions are diluted isothermally to 0.5 g/dl gives the heat of dilution at various temperatures. These calculated heats of dilution are compared with the experimental values obtained by carrying out the same dilutions in a microcalorimeter. The heat of dilution decreases linearly with β-casein concentration, but the extrapolated zero-concentration values of 65.8 ± 1.6 kJ mol?1 is the same as the Van't Hoff enthalpy. This agreement in the enthalpy values indicates that the micellization of β-casein occurs cooperatively. The effect of modifying the hydrophobic/hydrophilic balance of the system on the micellization of β-casein A has been investigated. The hydrophobic interaction between the protein molecules is decreased by removing the three C-terminal residues (Ileu Ileu Val) with carboxypeptidase-A. This modification drastically reduces the ability of the β-casein molecule to form micelles. Substitution of 2H2O for H2O at constant temperature perturbs the monomer–micelle equilibrium in favor of micelles because of enhanced hydrophobic interactions in the former solvent. The results are consistent with β-casein micellization involving a delicate balance of the hydrophobic forces favoring aggregation and electrostatic forces opposing it.  相似文献   

5.
Rates of approach to equilibrium values of F ST /R ST at various mutation rates and using different mutation models (K-allele model KAM and stepwise model SMM) were analyzed numerically for the finite island model and the one-dimensional stepping stone models of migration, using simulation. In the island model of migration and the KAM mutation model, the rate of approach to the equilibrium F ST value was appreciably higher and the equilibrium value was almost twofold lower at μ (mutation rate) = m (migration rate) than at μ ≪ m. In the one-dimensional stepping stone model of migration and the KAM model of mutation, the mutation rate significantly affected both the rate of approaching F ST equilibrium and the equilibrium value. In both island and one-dimensional stepping stone models and SMM, R ST was not influenced by various mutation rates. The rate of approach to the equilibrium values of both F ST and R ST was lower for the stepping stone model than to the island model. R ST was rather resistant to deviations from the SMM mutation model. __________ Translated from Genetika, Vol. 41, No. 9, 2005, pp. 1283–1288. Original Russian Text Copyright ? 2005 by Efremov.  相似文献   

6.
H Hervet  C P Bean 《Biopolymers》1987,26(5):727-742
The electrophoretic mobility (μ) of DNA fragments from λ phage and ΦX 174, split by restriction enzyme to molecular lengths from 3 × 102 to 2.36 × 104 base pairs, has been investigated in 0.6–4% agarose gels at various field strengths, ionic strengths, and temperatures. As already observed, μ is seen to be very sensitive to the field, increasing with field strength. The sensitivity increases with the molecular length of the DNA and decreases at high gel concentration. Our data are in qualitative agreement with recent theoretical predictions that concern the influence of the electric field on electrophoretic mobility. Mobility data have been extrapolated to zero field. This enables a comparison of our experimental results with theoretical predictions on the dependence of μ on the molecular weight of the DNA fragments. Our data fit, quite closely, a reptation model, where the tube path is described as a semiflexible entity with a persistence length equal to the pore diameter. The influence of the agarose concentration and the ionic strength of the buffer on the two parameters of the model—intrinsic electrophoretic mobility (μ0) and the number of base pairs per element of the tube (g)—are well described by the model. The temperature dependence of the electrophoretic mobility, together with the influence of the agarose concentration on μ0, indicate that the hydrodynamic drag is the leading frictional force on the DNA molecules in the gel.  相似文献   

7.
This study aimed to estimate trophic discrimination factors (TDFs) and metabolic turnover rates of nitrogen and carbon stable isotopes in blood and muscle of the smallnose fanskate Sympterygia bonapartii by feeding six adult individuals, maintained in captivity, with a constant diet for 365 days. TDFs were estimated as the difference between δ13C or δ15N values of the food and the tissues of S. bonapartii after they had reached equilibrium with their diet. The duration of the experiment was enough to reach the equilibrium condition in blood for both elements (estimated time to reach 95% of turnover: C t95%blood = 150 days, N t95%blood = 290 days), whilst turnover rates could not be estimated for muscle because of variation among samples. Estimates of Δ13C and Δ15N values in blood and muscle using all individuals were Δ13Cblood = 1·7‰, Δ13Cmuscle = 1·3‰, Δ15Nblood = 2·5‰ and Δ15Nmuscle = 1·5‰, but there was evidence of differences of c.0·4‰ in the Δ13C values between sexes. The present values for TDFs and turnover rates constitute the first evidence for dietary switching in batoids based on long‐term controlled feeding experiments. Overall, the results showed that S. bonapartii has relatively low turnover rates and isotopic measurements would not track seasonal movements adequately. The estimated Δ13C values in S. bonapartii blood and muscle were similar to previous estimations for elasmobranchs and to generally accepted values in bony fishes (Δ13C = 1·5‰). For Δ15N, the results were similar to published reports for blood but smaller than reports for muscle and notably smaller than the typical values used to estimate trophic position (Δ15N c. 3·4‰). Thus, trophic position estimations for elasmobranchs based on typical Δ15N values could lead to underestimates of actual trophic positions. Finally, the evidence of differences in TDFs between sexes reveals a need for more targeted research.  相似文献   

8.
Incubation of spinach thylakoids with HgCl2 selectively destroys Fe–S center B (FB). The function of electron acceptors in FB-less PS I particles was studied by following the decay kinetics of P700+ at room temperature after multiple flash excitation in the absence of a terminal electron acceptor. In untreated particles, the decay kinetics of the signal after the first and the second flashes were very similar (t 1/22.5 ms), and were principally determined by the concentration of the artificial electron donor added. The decay after the third flash was fast (t 1/20.25 ms). In FB-less particles, although the decay after the first flash was slow, fast decay was observed already after the second flash. We conclude that in FB-less particles, electron transfer can proceed normally at room temperature from FX to FA and that the charge recombination between P700+ and FX -/A1 - predominated after the second excitation. The rate of this recombination process is not significantly affected by the destruction of FB. Even in the presence of 60% glycerol, FB-less particles can transfer electrons to FA at room temperature as efficiently as untreated particles.Abbreviations DCIP 2, 6-dichlorophenol indophenol - FA, FB, FX iron-sulfur center A, B and X, respectively - PMS phenazine methosulfate  相似文献   

9.
Function of Y in codon-anticodon interaction of tRNA Phe   总被引:7,自引:0,他引:7  
Molar association constants of binding oligonucleotides to the anticodon loops of (yeast) tRNAPhe, (yeast) tRNAHClPhe and (E. coli) tRNAFMet have been determined by equilibrium dialysis. From the temperature dependence of the molar association constants, ΔF, ΔH and ΔS of oligomer-anticodon loop interaction have been determined. The data indicate that the free energy change of codon-anticodon interaction is highly influenced by the presence of a modified purine (tRNAPhe), of an unmodified purine (tRNAFMet) or its absence (tRNAHClPhe). Excision of the modified purine Y in the anticodon loop of tRNAPhe results in a conformational change of the anticodon loop, which is discussed on the basis of the corresponding changes in ΔF, ΔH and ΔS.  相似文献   

10.
Fructose, glucose, and mannose were treated with subcritical aqueous ethanol for ethanol concentrations ranging from 0 to 80% (v/v) at 180–200 °C. The aldose–ketose isomerization was more favorable than ketose–aldose isomerization and glucose–mannose epimerization. The isomerization of the monosaccharides was promoted by the addition of ethanol. In particular, mannose was isomerized most easily to fructose in subcritical aqueous ethanol. The apparent equilibrium constants for the isomerizations of mannose to fructose, Keq,M→F, and glucose to fructose, Keq,G→F, were independent of ethanol concentration and increased with increasing temperature. Moreover, the Keq,M→F value was much larger than the Keq,G→F value. The enthalpies for the isomerization of mannose to fructose, ΔHM→F, and glucose to fructose, ΔHG→F, were estimated to be 18 and 24 kJ/mol, respectively, according to van’t Hoff equation. Subcritical aqueous ethanol can be used to produce fructose from glucose and mannose efficiently.  相似文献   

11.
A mathematical model for single and multi step deep-jet bioreactors is presented. A stagewise approach based on macroscopic mechanistic model which divides the reactor into compartments with good quality of mixing and plug flow regions (macromixer), was used. For the mathematical representation of this model a system of differential equations, describing the concentration of tracer in structural elements based on mass balance, and the Runge-Kutta-Fehlberg numerical method of integration, was applied. The mixing time in a 300 dm3 tank was determined by conductivity method with NaCl as tracer.List of Symbols V g dm3 total volume of liquid - V 1; V 6 dm3 volumes of ideally mixed compartments in the vessel - V 2; V 7 dm3 volumes of macromixer in the inner circulation flows - V 3; V 9 dm3 volumes of liquid phase in the pump - V 4; V 8 dm3 volumes of liquid phase in the pipe between the vessel and the pump - V 5; V 10 dm3 volumes of liquid phase in pipes between the pump and the air input system, including falling jet - F E; F E,1; F E,2 dm3/s the inner volumetric circulation flow rates accross the macromixers - F E,3; F E,4 dm3/s exchanges volumetric flow rates between two ideally mixed compartments in the vessel - F cir; F 1,cir; F 2,cir dm3/s external volumetric circulation flow rates (pumping capacity) - t A s time interval of puls application - t AA s time point of impuls application related to the free chosen point of simulation - t end s end time of simulation - F qu g2/dm6 sum of quadratic error - C *,* kg/m3 concentration of the tracer in the indicated compartment - C 0 kg/m3 concentration of the tracer before the injection - C t kg/m3 concentration of the tracer at the indicated time - C kg/m3 theoretical concentration of full mixed tracer - i index of an arbitrary tank - C sim kg/m3 calculated concentration of the tracer by numerical integration method  相似文献   

12.
An amorphous solid of cyclomaltoheptaose (β-cyclodextrin, β-CD) was formed by milling its crystalline form using a high-energy planetary mill at room temperature. The glass transition of this amorphous solid was found to occur above the thermal degradation point of the material preventing its direct observation and thus its full characterization. The corresponding glass transition temperature (Tg) and the ΔCp at Tg have, however, been estimated by extrapolation of Tg and ΔCp of closely related amorphous compounds. These compounds include methylated β-CD with different degrees of substitution and molecular alloys obtained by co-milling β-CD and methylated β-CD (DS 1.8) at different ratios. The physical characterization of the amorphous states have been performed by powder X-ray diffraction and differential scanning calorimetry, while the chemical integrity of β-CD upon milling was checked by NMR spectroscopy and mass spectrometry.  相似文献   

13.
A strict analytical theory has been developed describing the behavior of a model lattice polymer chain of arbitrary stiffness in a slitlike pore at polymer–adsorbent interaction energies –ε. The thermodynamic characteristics of the system were calculated. It was shown that the transition of the macromolecule from the solution volume inside a pore occurs by the first-order phase transition with evolution of latent heat of adsorption. The transition point –ε = –εc is determined by the chain stiffness and is independent of the pore width D. It is shown that in the precritical range, –ε < –εc, the free energy ΔF of the macromolecules in the pores is adequately described by the universal dependence ΔF = ΔF(D*/A), where D* is some effective pore width depending on the value of –ε, and A is the length of the Kuhn segment. At high attraction energies, –ε ? –εc, the macromolecules are bonded to the pore walls by a great number of units and their free energy depends only on –ε and the chain stiffness, ΔF = ΔF(A, ε). Close to the critical energy –ε ? –εc (transition range), ΔF is determined by both the stiffness of the macromolecule and the pore width D: ΔFA2D?1 for fairly high values of A and D. The possibilities of using porous media as protein stabilizers are discussed, and the value of the stabilizing effect depending on the chain stiffness is estimated.  相似文献   

14.
The effect of temperature, light-spectrum, desiccation and salinity gradients on the photosynthesis of a Japanese subtidal brown alga, Sargassum macrocarpum (Fucales), was determined using a pulse amplitude modulation-chlorophyll fluorometer and dissolved oxygen sensors. Temperature responses of the maximum (Fv/Fm in darkness) and effective (ΔF/Fm at 50 μmol photons m−2 s−1; = ΦPSII) quantum yields during 6-day culture (4–36°C) remained high at 12–28°C, but decreased at higher temperatures. Nevertheless, ΔF/Fm also dropped at temperatures below 8°C, suggesting light sensitivity under chilling temperatures because Fv/Fm remained high. Photosynthesis–irradiance responses at 24°C under red (660 nm), green (525 nm), blue (450 nm) and white light (metal halide lamp) showed that maximum net photosynthesis under blue and white light was greater than under red and green light, indicating the sensitivity and photosynthetic availability of blue light in the subtidal light environment. In the desiccation experiment, samples under aerial exposure of up to 8 h under dim-light at 24°C and 50% humidity showed that ΔF/Fm quickly declined after more than 45 min of emersion; furthermore, ΔF/Fm also failed to recover to initial levels even after 1 day of rehydration in seawater. Under the emersion state, the ΔF/Fm remained high when the relative water content (RWC) was greater than 50%; in contrast, it quickly dropped when the RWC was less than 50%. When the RWC was reduced below 50%, ΔF/Fm did not return to initial levels, regardless of subsequent re-hydration, suggesting a low capacity of photosynthesis to recover from desiccation. The stenohaline response of photosynthesis under 3-day culture is evident, given that ΔF/Fm declined when salinity was beyond 20–40 psu. Adaptation to subtidal environments in temperate waters of Japan can be linked to these traits.  相似文献   

15.
Summary Cell recovery by means of continuous flotation of the Hansenula polymorpha cultivation medium without additives was investigated as a function of the cultivation conditions as well as of the flotation equipment construction and flotation operational parameters. The cell enrichment and separation is improved at high liquid residence times, high aeration rates, small bubble sizes, increasing height of the aerated column, and diameter of the foam column. Increasing cell age and cultivation with nitrogen limitation reduce the cell separation.Symbols CP cell mass concentration in medium g·l–1 - CR cell mass concentration in residue g·l–1 - CS cell mass concentration in foam liquid g·l–1 - V equilibrium foam volume cm3 - V gas flow rate through the aerated liquid column cm3·s–1 - VF feed rate to the flotation column ml/min - 1 V S/V foaminess s - mean liquid residence time in the column s  相似文献   

16.
A series of new alkyl and alkoxide (FPNP)Pd complexes have been synthesized. The alkyls and alkoxides containing β-hydrogens display remarkable thermal stability. Thermal decomposition of (FPNP)PdOEt is very slow in pure C6D6 but is accelerated by the addition of EtOH co-solvent. It is proposed that the β-hydrogen elimination from (FPNP)Pd-OCH2R occurs via dissociation of the alkoxide anion.  相似文献   

17.
An investigation was performed into the operation of an integrated system for continuous production and product recovery of solvents (acetone-butanol-ethanol) from the ABE fermentation process. Cells of Clostridium acetobutylicum were immobilized by adsorption onto bonechar, and used in a fluidized bed reactor for continuous solvent production from whey permeate. The reactor effluent was stripped of the solvents using nitrogen gas, and was recycled to the reactor. This relieved product inhibition and allowed further sugar utilization. At a dilution rate of 1.37 h–1 a reactor productivity of 5.1 kg/(m3 · h) was achieved. The solvents in the stripping gas were condensed to give a solution of 53.7 kg/m3. This system has the advantages of relieving product inhibition, and providing a more concentrated solution for recovery by distillation. Residual sugar and non-volatile reaction intermediates are not removed by gas stripping and this contributes to high solvent yields.List of Symbols C kg/m3 Lactose concentration in reactor effluent - C b kg/m3 Lactose concentration in bleed stream - C c kg/m3 Lactose concentration in whey permeate feed - C i kg/m3 Lactose concentration at reactor inlet - C p kg/m3 Lactose concentration in condensed solvent stream (=0) - C r kg/m3 Lactose concentration in recycle line (C b=C r) - C kg/h Amount of lactose utilized during certain time period - D h1 Dilution rate of reactor, F i/D=F/D - F dm3/h, m3/h F i = Rate of feed flow to the reactor - F b dm 3/h, m3/h Rate of bleed - F c dm3/h, m3/h Rate of feed of whey permeate solution - F p dm3/h, m3/h Rate of concentrated product removal - F r dm3/h, m3/h Rate of recycle of stripped effluent to the reactor - P l % Percent lactose utilization - R l kg/(m3 · h) Overall lactose utilization rate - R p kg/(m3 · h) Overall reactor (solvent) productivity - R sl kg/h Rate of solvent loss - S kg/m3 Solvent concentration in reactor effluent - S b kg/m3 Solvent concentration in bleed - S c kg/m3 0; Solvent concentration in concentrated whey permeate solution - S i kg/m3 Solvent concentration at inlet of reactor - S p kg/m3 Solvent concentration in concentrated product stream - S r kg/m3 Solvent concentration in stripped effluent, S r=Sb - S kg/h Amount of solvent produced from C amount of lactose in a particular time - ds/dt kg/(m3 · h) Rate of accumulation of solvents in the stripper - t h Time - V dm3, m3 Total reactor volume - V 1 dm3, m3 Liquid volume in stripper - Y P/S Solvent yield  相似文献   

18.
The susceptibility to photoinhibition of tree species from three different successional stages were examined using chlorophyll fluorescence and gas exchange techniques. The three deciduous broadleaf tree species were Betula platyphylla var. japonica, pioneer and early successional, Quercus mongolica, intermediate shade‐tolerant and mid‐successional, and Acer mono, shade‐tolerant and late successional. Tree seedlings were raised under three light regimes: full sunlight (open), 10% full sun, and 5% full sun. Susceptibility to photoinhibition was assessed on the basis of the recovery kinetics of the ratio of vaviable to maximum fluorescence (Fv/Fm) of detached leaf discs exposed to about 2000 μmol m?1 s?1 photon flux density (PFD) for 2 h under controlled conditions (25 to 28 °C, fully hydrated). Differences in susceptibility to photodamage among species were not significant in the open and 10% full sun treatments. But in 5% full sun, B. platyphylla sustained a significantly greater photodamage than other species, probably associated with having the lowest photosynthetic capacity indicated by light‐saturated photosynthetic rate (B. platyphylla, 9·87, 5·85 and 2·82; Q. mongolica, 8·05, 6·28 and 4·41; A. mono, 7·93, 6·11 and 5·08 μmol CO2 m?1 s?1for open, 10% and 5% full sun, respectively). To simulate a gap formation and assess its complex effects including high temperature and water stress in addition to strong light on the susceptibility to photoinhibition, we examined photoinhibition in the field by means of monitoring ΔF/Fm on the first day of transfer to natural daylight. Compared with ΔF/Fm in AM, the lower ΔF/Fm in PM responding to lower PFD following high PFD around noon indicated that photoinhibition occurred in plants grown in 10 and 5% full sun. The diurnal changes of ΔF/Fm showed that Q. mongolica grown in 5% full sun was less susceptible to photoinhibition than A. mono although they showed little differences both in photosynthetic capacity in intact leaves and susceptibility to photoinhibition based on leaf disc measurements. These results suggest that shade‐grown Q. mongolica had a higher tolerance for additional stresses such as high temperature and water stress in the field, possibly due to their lower plasticity in leaf anatomy to low light environment.  相似文献   

19.
We present evidence that plant growth at elevated atmospheric CO2 increases the high‐temperature tolerance of photosynthesis in a wide variety of plant species under both greenhouse and field conditions. We grew plants at ambient CO2 (~ 360 μ mol mol ? 1) and elevated CO2 (550–1000 μ mol mol ? 1) in three separate growth facilities, including the Nevada Desert Free‐Air Carbon Dioxide Enrichment (FACE) facility. Excised leaves from both the ambient and elevated CO2 treatments were exposed to temperatures ranging from 28 to 48 °C. In more than half the species examined (4 of 7, 3 of 5, and 3 of 5 species in the three facilities), leaves from elevated CO2‐grown plants maintained PSII efficiency (Fv/Fm) to significantly higher temperatures than ambient‐grown leaves. This enhanced PSII thermotolerance was found in both woody and herbaceous species and in both monocots and dicots. Detailed experiments conducted with Cucumis sativus showed that the greater Fv/Fm in elevated versus ambient CO2‐grown leaves following heat stress was due to both a higher Fm and a lower Fo, and that Fv/Fm differences between elevated and ambient CO2‐grown leaves persisted for at least 20 h following heat shock. Cucumis sativus leaves from elevated CO2‐grown plants had a critical temperature for the rapid rise in Fo that averaged 2·9 °C higher than leaves from ambient CO2‐grown plants, and maintained a higher maximal rate of net CO2 assimilation following heat shock. Given that photosynthesis is considered to be the physiological process most sensitive to high‐temperature damage and that rising atmospheric CO2 content will drive temperature increases in many already stressful environments, this CO2‐induced increase in plant high‐temperature tolerance may have a substantial impact on both the productivity and distribution of many plant species in the 21st century.  相似文献   

20.
The synthesis of new β-diketonato rhodium(I) complexes of the type [Rh(FcCOCHCOR)(CO)2] and [Rh(FcCOCHCOR)(CO)(PPh3)] with Fc=ferrocenyl and R=Fc, C6H5, CH3 and CF3 are described. 1H, 13C and 31P NMR data showed that for each of the non-symmetric β-diketonato mono-carbonyl rhodium(I) complexes, two isomers exist in solution. The equilibrium constant, Kc, which relates these two isomers in an equilibrium reaction, are concentration independent but temperature and solvent dependent. ΔrG, ΔrH and ΔrS values for this equilibrium have been determined and a linear relationship between solvent polarity on the Dimroth scale and Kc exists. The relationship between RhP bond lengths, d(RhP), and 31P NMR peak positions as well as coupling constants 1J(31P103Rh) has been quantified to allow calculation of approximate d(RhP) values. Variations in d(RhP) for [Rh(RCOCHCOR′)(CO)(PPh3)] complexes have also been related to the group electronegativities (Gordy scale) of the terminal β-diketonato R groups trans to PPh3. A measure of the electron density on the rhodium centre of [Rh(RCOCHCOR′)(CO)(PPh3)] may be expressed in terms of the IR carbonyl stretching wave number, ν(CO), the sum of the group electronegativities of the R and R′ groups, (χR+χR′), or the observed pKa values of the free β-diketones RCOCH2COR. An empirical relationship between ν(CO) and either pKa or (χR+χR′) has also been quantified.  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