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 46 毫秒
1.
Gas exchange characteristics of three major Louisiana Mississippi River deltaic plain marsh species, Spartina patens (Ait.) Muhl., Spartina altemiflora Lois., and Panicum hemitomon Shult., was studied under controlled environment conditions. The optimum temperature for maximum photosynthesis was ≈ 36 °C for S. patens, 27 °C for S. alterniflora, and 28 °C for rP. hemitomon. Net photosynthesis rates at optimum temperature averaged 20.1 μmol · mt-2 · st-1 in S. patens, 22.8 μmol · m−2 · s−1 in S. alterniflora, and 11.4 μmol · m−2 · s−1 in P. hemitomon. Photosynthetic light saturation occurred ≈720, 530, and 750 μmol · m−2 · s−1 in S. patens, S. alterniflora, and P. hemitomon, respectively. Only S. patens had a midday depression of stomatal conductance, but net photosynthesis was not reduced by the depression. Maximum stomatal conductances were 285 mmol · m−2 · s−1 in S. patens, 238 mmol · m−2 · s−1 in S. alterniflora, and 335 mmol · m−2 · s−1 in P. hemitomon. In contract, net photosynthesis values were lower in P. hemitomon compared with the Spartina species, indicating a greater degree of water use efficiency of photosynthesis for both Spartina species.  相似文献   

2.
Classical NaCa exchange models are based on a symmetric carrier system where Na and Ca competing from the same site, can produce net movement of the other against its electrochemical gradient. We have explored this symmetric assumption by studying the Cao and Nao-dependent Na efflux in dialyzed squid axons in which proper control of both external and internal medium was achieved. The results show: (1) In axons dialyzed without Cai and ATP, Cao-dependent Na efflux cannot be detected even in the absence of Nao. Under these conditions, the level of Na efflux (1 pmol · cm−2 · s−1) is close to that predicted by an electrical ‘leak’. (2) In axons dialyzed with Cai (100 μM) and without ATP, Na efflux measured in 440 mM Nao, is about 4–5 pmol · cm−2 · s−1 and rather insensitive to Cao between 0 and 10 mM. However, in the absence of Nao, a Cao-dependent Na efflux is observed similar in magnitude to that found in the presence of external Na. (3) In the presence of both Cai and ATP, Na efflux into artificial sea-water (440 mM Na, 10 mM Ca) is 18 pmol · cm−2 · s−1. In the absence of Nao the efflux of Na is 7.5 pmol · cm−2 · s−1. In the absence of both Nao and Cao the efflux is close to ‘leak’. With full Nao but no Cao, the Na efflux average 12.6 pmol · cm−2 · s−1. These results indicate a marked asymmetry in the modus operandi of the NaCa exchange system with respect to Cai and ATP. These two substrates are required from the cis side to promote Cao-dependent Na efflux (reversal NaCa exchange).  相似文献   

3.
We have used inelastic laser light scattering to study the kinetics of the spontaneous assembly of heads and tails of bacteriophage T4D to form noninfectious tail fiberless particles. For interpretation of the kinetics, it was first necessary to determine the physical properties of the strongly scattering phage parts. For heads, these are D20,w = 3.60 × 10−8cm2/s, 820,w = 1025 S, M = 1.76 × 108. For tail fiberless particles, D20,w = 3.14 × 10−8cm2/s, 820,w = 968 S, and M = 1.95 × 108. The kinetics of the head-tail joining process was followed by measuring the time variation of the homodyne scattering autocorrelation function. This was interpreted as a sum of exponentials whose decay constants were known from the scattering angle and the diffusion coefficients, and whose amplitudes were related to the concentrations of reactants and products. Scattering experiments at 22 °C gave a bimolecular rate constant of 1.02 × 107m−1 s−1, while infectivity assays at 30 °C gave a rate constant of 1.28 × 107. Adjustment of both rate constants to 20 °C, assuming diffusion controlled reaction, gave 0.97 × 107 and 0.98 × 107m−1 s−1, respectively. This rate is about 1500 that predicted by Smoluchowski theory for a diffusion controlled reaction between two spherical particles; the discrepancy is largely explicable from orientational factors.  相似文献   

4.
The efficacy of visual and non-visual feeding among pelagic striped bass Morone saxatilis larvae adapted to a turbid estuary was determined in the laboratory in clear water. Capture of Artemia salina (density 100 l1) was significantly affected by the interaction between age of larvae (range: 8–25 days post-hatch, dph) and light intensity (range: 0–10·6 μmol s1 m2 at the water surface). Visual feeding by larvae aged 9–11 dph was highest in dim light (0·086–0·79 μmol s1 m2), with fish capturing up to 5 prey larva1 h1. As the larvae grew, prey capture in brighter light improved, associated with an increasing proportion of twin cone photoreceptors and improving ability of the retina to light- and dark-adapt. By age >22 dph, mean prey capture was greatest at highest light intensities (0·79 and 10·6 μmol s1 m2) exceeding 100 prey larva1 h1. Incidence of feeding larvae generally improved as the larvae grew, reaching >80% in all light intensities from 16 dph onwards. The lower threshold for visual feeding, between 0·0084 and 0·03 μmol s1 m2, remained constant as the larvae grew, despite an increasing density of rod photoreceptors. Below this threshold, non-visual feeding was evident at a low rate (<6 prey larva1 h1) that was independent of larval age.  相似文献   

5.
Tissue kallikrein may play a role in processing precursor polypeptide hormones. We investigated whether hydrolysis of natural enkephalin precursors, peptide F and bovine adrenal medulla docosapeptide (BAM-22P), by hog pancreatic kallikrein is consistent with this concept. Incubation of peptide F with this tissue kallikrein resulted in the release of Met5-enkephalin and Met5-Lys6-enkephalin. Met5-Lys6-enkephalin was the main peptide released, indicating that the major cleavage site was between two lysine residues. At 37°C and pH 8.5, the KM values for formation of Met5-enkephalin and Met5-Lys6-enkephalin were 129 and 191 μM, respectively. Corresponding kcat values were 0.001 and 0.03 s−1 and kcat/KM ratios were 8 and 1.6·102 M−1 · s−1, respectively. Cleavage of peptide F at acidic pH (5.5) was negligible. When BAM-22P was used as a substrate, Met5-Arg6-enkephalin was released, thus indicating cleavage between two arginine residues. At pH 8.5, KM was 64 μM, kcat was 4.5 s−1, and the kcat/KM ratio was 7 · 104 M−1 · s−1. At 5.5, the pH of the secretory granules, KM, kcat and kcat/KM were 184 μM, 1.9 s−1 and 104 M−1 · s−1, respectively. It is unlikely that peptide F could be a substrate for kallikrein in vivo; however, tissue kallikrein could aid in processing proenkephalin precursors such as BAM-22P by cleaving Arg-Arg peptide bonds.  相似文献   

6.
The longitudinal diffusion of a homologous series of monoamides through lecithin-water lamellar phases with aqueous channel widths of 16–27 Å has been studied. The diffusion coefficients relative to water of the hydrophilic amides, formamide and acetamide, depend logarithmically on solute molar volume, as previously demonstrated in human red cells. Aqueous diffusion of amides in red-cell membranes is similar to that in a lecithin-water phase of aqueous channel width less than 16 Å, the smallest channel width used. Partition coefficients of the lipophilic amides, valeramide and isovaleramide, between lecithin vesicles and water are 1.64 and 1.15 at 20 °C. These data enabled us to compute a valeramide diffusion coefficient of 6.5 · 10−7cm2 · s−1 at 20 °C in the lipid region of a lamellar phase containing 30% water about one order of magnitude greater than the diffusion coefficient of spin-labelled analogs of phosphatidylcholine. The discrimination between the permeability coefficients of valeramide and isovaleramide is more than twice as great in the human red cell as between lipid diffusion coefficients in a phase containing 8% water. This suggests that the lipid region of the human red cell is more highly organized than lipid in the lecithin-water lamellar phase.  相似文献   

7.
The free solution mobility of DNA has been measured by capillary electrophoresis in the two buffers most commonly used for DNA gel electrophoresis, Tris-borate-EDTA (TBE) and Tris-acetate-EDTA (TAE). The capillaries were coated with polymers of either of two novel acrylamide monomers, N-acryloylaminoethoxyethanol or N-acryloylaminopropanol, both of which are stable at basic pH and effectively eliminate the electroendosmotic mobility due to the capillary walls. The free solution mobility of DNA in TAE buffer was found to be (3.75 ± 0.04) × 10−4 cm2 V−1 s−1 at 25°C, independent of DNA concentration, sample size, electric field strength, and capillary coating, and in good agreement with other values in the literature. The free solution mobility was independent of DNA molecular weight from ∼ 400 base pairs to 48.5 kilobase pairs, but decreased monotonically with decreasing molecular weight for smaller fragments. Surprisingly, the free solution mobility of DNA in TBE buffer was found to be (4.5 ± 0.1) × 10−4 cm2 V−1 s−1, about 20% larger than observed in TAE buffer, presumably because of the formation of nonspecific borate-deoxyribose complexes. © 1997 John Wiley & Sons, Inc. Biopoly 42: 687–703, 1997  相似文献   

8.
The electrophoretic mobility of L5178Y cells in 0.0145 M NaCl, 4.5% sorbitol, 0.6 mM NaHCO3, pH 7.2, at 25°C was — 1.78 μ·s?1·V?1·cm?1 while that of an L-asparaginase resistant subline, L5178Y/ASN, was — 1.11 μm·s?1·V?1·cm?1. Both cell lines were characterized by terminal sialic acid residues on their surfaces. Treatment of L5178Y cells for 90 min with 10 units of L-asparaginase per ml in saline decreased the electrophoretic mobility of the cells to — 1.65 μm·s?1·V?1·cm?1 while treatment in Fischer's medium decreased the mobility to — 1.25 μm·s?1·V?1·cm?1; neither treatment had a significant effect on the L5178Y/ASN electrophoretic mobility. The results suggest that L-asparaginase has an immediate and specific effect on synthesis of cell surface asparaginyl glycoproteins.  相似文献   

9.
The kinetics of the chlorophyll fluorescence rise induced by adding 20 mM MgCl2 to a suspension of isolated pea chloroplasts treated with 3-(3,4-dichlorophenyl)-1, 1-dimethylurea (DCMU) have been examined experimentally and theoretically as a function of temperature. The application of similarity arguments and particle aggregation theory to the experimental results suggests that at the first approximation, the salt-induced time-dependent fluorescence changes may be described by the diffusion-controlled lateral movement of Photosystem II pigment-protein complexes. From an analysis of the temperature dependence of the fluorescence changes, estimates obtained for the lateral diffusion coefficients were 1.85 · 10?12–3.08 · 10?11 cm2/s over the temperature range 10°C ? T?30°C.  相似文献   

10.
At pH 8.0 and 25°C α1-proteinase inhibitor and α2-macroglobulin bind human pancreatic elastase with rate constants of 4.7·105 M−1·s−1 and 6.4·106 M−1·s−1, respectively. The corresponding delay times of elastase inhibition in plasma are 0.4 s and 0.2 s, respectively, indicating that both inhibitors may act as physiological antielastases. Elastin impairs the elastase inhibitory capacity of α1-proteinase inhibitor and α2-macroglobulin. In presence of human elastin, the former behaves like a slow-binding elastase inhibitor, with a rate constant of about 260 M−1·s−1. In contrast, α2-macroglobulin is a fast-binding inhibitor of elastin-bound elastase, but only one of its two sites is functioning in presence of elastin.  相似文献   

11.
A laboratory-made sample of the polysaccharide xylinan (acetan) has been further characterized with respect to (i) purity, (ii) molar mass and polydispersity, and (iii) gross conformation by a combination of hydrodynamic measurements (sedimentation velocity and equilibrium analytical ultracentrifugation, viscometry, and dynamic light scattering) in aqueous NaCl (I = 0.10 mol·L−1). Sedimentation velocity diagrams recorded using Schlieren optics revealed highly pure material sedimenting as a single boundary [so20.w = 9.5 ± 0.7) S; ks = (273 ± 112) mL/g]. The hypersharp nature of these boundaries is symptomatic of a polydisperse and highly nonideal (in the thermodynamic sense) system. Low speed sedimentation equilibrium in the analytical ultracentrifuge using Rayleigh interference optics and two different types of extrapolation procedure (involving point and whole-cell molar masses) gave a weight average molar mass Mw of (2.5 ± 0.5) × 10−6 g·mol−1 and also a second virial coefficient, B = (2.8 ± 0.7) × 10−4 mL·mol·g−2, both values in good agreement with those from light scattering-based procedures (Part II of this series). A dynamic Zimm plot from dynamic light scattering measurements gave a z-average translational diffusion coefficient Do20.w = (3.02 ± 0.05) × 10−8 cm2·s−1 and the concentration-dependence parameter kD = (370 ± 15) mL/g. Combination of so20.w with Do20.w via the Svedberg equation gave another estimate for Mw of ≅ 2.4 × 106 g/mol, again in good agreement. Both the Wales-van Holde ratio (ks/[η]) ≅ 0.4 (with [η] = (760 ± 77) mL/g) and the ρ-parameter (ratio of the radius of gyration from static light scattering to the hydrodynamic radius from dynamic light scattering) as ρ > 2.0 all indicate an extended conformation for the macromolecules in solution. These findings, plus Rinde-type simulations of the sedimentation equilibrium data are all consistent with the interpretation in terms of a unimodal wormlike coil model performed earlier. © 1996 John Wiley & Sons, Inc.  相似文献   

12.
The reaction between ligninase and hydrogen peroxide yielding Compound I has been investigated using a stopped-flow rapid-scan spectrophotometer. The optical absorption spectrum of Compound I appears different to that reported by Andrawis, A. et al. (1987) and Renganathan, V. and Gold, M.H. (1986), in that the Soret-maximum is at 401 nm rather than 408 nm. The second-order rate constant (4.2·105 M−1·s−1) for the formation of Compound I was independent of pH (pH 3.0–6.0). In the absence of external electron donors, Compound I decayed to Compound II with a half-life of 5–10 s at pH 3.1. The rate of this reaction was not affected by the H2O2 concentration used. In the presence of either veratryl alcohol or ferrocyanide, Compound II was rapidly generated. With ferrocyanide, the second-order rate constant increased from 1.9·104 M−1·s−1 to 6.8·106 M−1·s−1 when the pH was lowered from 6.0 to 3.1. With veratryl alcohol as an electron donor, the second-order rate constant for the formation of Compound II increased from 7.0·103 M−1·s−1 at pH 6.0 to 1.0·105 M−1·s−1 at pH 4.5. At lower pH values the rate of Compound II formation no longer followed an exponential relationship and the steady-state spectral properties differed to those recorded in the presence of ferrocyanide. Our data support a model of enzyme catalysis in which veratryl alcohol is oxidized in one-electron steps and strengthen the view that veratryl alcohol oxidation involves a substrate-modified Compound II intermediate which is rapidly reduced to the native enzyme.  相似文献   

13.
β-Mannanase is the key enzyme in the hydrolysis of mannan which has been widely applied in diverse industrial fields such as biobleaching pulps, food and feed industry, bioethanol and pharmaceutical applications. In this study, a novel GH5 family β-mannanase gene (LrMan5B) with 381 amino acid residues was identified from Lichtheimia ramosa, and highly expressed in Pichia pastoris X33. The amino acid sequence shares the highest identity (64%) with the β-mannanase from Rhizomucor miehei. Purified recombinant LrMan5B showed the optimal activity at pH 5.0 and 65 °C. It had broad-range pH stability (retaining >65% activity after incubation at pH 3.0–8.0 at 37 °C for 24 h) and was highly thermostable (retaining >80% activity after incubation at 60 °C for 30 min). LrMan5B displayed the highest catalytic efficiency for locust bean gum and the kcat/Km value was 1357.47 mL·mg−1·s−1, followed by guar gum (512.82 mL·mg−1·s−1), konjac glucomannan (454.21 mL·mg−1·s−1), and palm kernel meal (137.00 mL·mg−1·s−1). In order to evaluate the synergistic effect of LrMan5B and α-galactosidase LrAgal36A from L. ramosa, LrAgal36A was supplemented to hydrolyze palm kernel meal with LrMan5B together, showing that the reducing sugar release significantly increased by 21% (compared with the sum of that by hydrolysis of single Lrman5B or LrAgal36A). Due to its favorable enzymatic properties, LrMan5B might own potential applications in the area of food and feed processing.  相似文献   

14.
《Bioorganic chemistry》1987,15(2):100-108
Nonenzymatic rates of hydrolytic deamination of adenosine and cytidine by acids and bases analogous to side chains of naturally occurring amino acids are compared with the rates of uncatalyzed deamination in water and with the rates of the hydroxide- and hydrogen ion-catalyzed reactions. For adenosine, hydroxide ion is an effective catalyst, with a second-order rate constant of 7.5 × 10−6 m−1 s−1 at 85°C and an energy of activation of 19.9 kcal/mol. Acid-catalyzed deamination of adenine proceeds with a second-order rate constant of 1.5 × 10−6 m−1 s−1 at 85°C. At concentrations of 1 m and at pH values corresponding to their respective pKa values, dimethylamine, acetate, selenide, imidazole, phosphate, and zinc(II) do not enhance the rate of deamination of adenosine beyond that observed in water, and 2-mercaptoethanol produces only a modest rate enhancement. The uncatalyzed rate of adenosine deamination in water is 8.6 × 10−9 s−1 at 85°C: extrapolation to 37°C and comparison with kcat for rat hepatoma adenosine deaminase yield a rate enhancement by the enzyme of approximately 2 × 1012-fold. 1,6-Dimethyladenosine, the conjugate acid of which has a pKa value much higher than that of adenosine, is not readily deaminated, suggesting that the uncatalyzed deamination of adenosine does not proceed by hydroxide ion attack on the rare protonated form of adenosine, but rather by attack on the neutral species. Deamination of cytidine is catalyzed most effectively by hydroxide ion, with a second-order rate constant of 4.5 × 10−4 m−1 s−1 at 85°C and an energy of activation of 28.5 kcal/mol. The uncatalyzed rate of deamination of cytidine in water, which also exhibits an energy of activation of 28.5 kcal/mol, is 8.8 × 10−8 s−1 at 85°C. Comparison of the rate extrapolated to 25°C with kcat for bacterial cytidine deaminase gives a rate enhancement for the enzyme of 4 × 1011-fold. The C-5 proton of the pyrimidine ring of cytidine does not exchange with solvent during alkaline hydrolysis, suggesting that deamination under these conditions does not involve prior addition of water across the 5,6 double bond.  相似文献   

15.
The relationship between leaf resistance to water vapour diffusion and each of the factors leaf water potential, light intensity and leaf temperature was determined for leaves on seedling apple trees (Malus sylvestris Mill. cv. Granny Smith) in the laboratory. Leaf cuticular resistance was also determined and transpiration was measured on attached leaves for a range of conditions. Leaf resistance was shown to be independent of water potential until potential fell below — 19 bars after which leaf resistance increased rapidly. Exposure of leaves to CO2-free air extended the range for which resistance was independent of water potential to — 30 bars. The light requirement for minimum leaf resistance was 10 to 20 W m?2 and at light intensities exceeding these, leaf resistance was unaffected by light intensity. Optimum leaf temperature for minimum diffusion resistance was 23 ± 2°C. The rate of change measured in leaf resistance in leaves given a sudden change in leaf temperature increased as the magnitude of the temperature change increased. For a sudden change of 1°C in leaf temperature, diffusion resistance changed at a rate of 0.01 s cm?1 min?1 whilst for a 9°C leaf temperature change, diffusion resistance changed at a rate of 0.1 s cm?1 min?1. Cuticular resistance of these leaves was 125 s cm?1 which is very high compared with resistances for open stomata of 1.5 to 4 s cm?1 and 30 to 35 s cm?1 for stomata closed in the dark. Transpiration was measured in attached apple leaves enclosed in a leaf chamber and exposed to a range of conditions of leaf temperature and ambient water vapour density. Peak transpiration of approximately 5 × 10?6 g cm?2 s?1 occurred at a vapour density gradient from the leaf to the air of 12 to 14 g m?3 after which transpiration declined due presumably to increased stomatal resistance. Leaves in CO2-free air attained a peak transpiration of 11 × 10?6 g cm?2 s?1 due to lower values of leaf resistance in CO2 free air. Transpiration then declined in these leaves due to development of an internal leaf resistance (of up to 2 s cm?1). The internal resistance was masked in leaves at normal CO2 concentrations by the increase in stomatal resistance.  相似文献   

16.
Pulse radiolytic studies of α-tocopherol (αTH) oxidation-reduction processes were carried out with low doses (5 Gy) of high-energy electrons in O2−, N2−, and air-saturated ethanolic solutions. Depending on the concentration of oxygen in solution, two different radicals, A· and B·, were observed. The first, A·, was obtained under N2 and results from aTH reaction with solvated electron (kaTH+csolv = 3.4 × 108 mol−1 liter s−1) and with H3C-ĊH-OH, (R·) (kaTH + R· = 5 × 105 mol−1 liter s−1). B·, observed under O2, is produced by αTH reaction with RO2 peroxyl radicals (kaTH + RO2. = 9.5 × 104 mol−1 liter s−1).  相似文献   

17.
《BBA》1985,807(3):221-229
Bacteriochlorophyll (BChl) luminescence lifetimes (τ) were measured in purple bacteria Rhodospirillum rubrum and Rhodopseudomonas sphaeroides at low-excitation pulse energy with the use of a picosecond luminescence spectrochronograph of high sensitivity and high time-resolution. Average high-frequency excitation light density was changed from about 1 · 1013 photons · cm−2 · s−1 up to 1 · 1017 photons · cm−2. s−1. Maximal energy density in a single pulse was in the range 10−14–10−10 J/cm2, which completely rules out nonlinear exciton interactions. In this range τ increased as a function of excitation light density from about 60 ps to 210 ps. Luminescence yield (ø) for the bacteria investigated measured under continuous or picosecond excitation changed in a similar manner as τ. The luminescence increase was shown to accompany the conversion of the reaction centers to the closed, photooxidized state. Luminescence decay of R. rubrum and Rps. sphaeroides chromatophores without any chemical additions was well approximated by a single exponential component both at low and at saturating intensities of exciting light. The time necessary for the primary charge separation to occur was shown to be 60 ± 10 ps. The pairwise jump-time of excitation-energy transfer, as well as excitation-diffusion characteristics were estimated from these data. On the basis of life-time measurements in the state of active photosynthesis, the quantum yield of the primary charge separation in the reaction centers was estimated to be equal to 0.95 ± 0.02. In intact cells as well as in chromatophores in the presence of reducing agents, a nanosecond component of emission decay was also observed. The relative amplitude of this component, being several percent of the picosecond one at low-excitation intensity levels, increased (2–3)-times with excitation density. Its life-time was estimated to be 3 ± 1 ns. The nanosecond component appeared only under conditions when a part of the reaction centers were converted to the closed state PQ.  相似文献   

18.
  • 1.1. The diffusional water permeability (Pd) of rabbit red blood cell (RBC) membrane has been monitored by a doping nuclear magnetic resonance (NMR) technique on control cells and following inhibition with p-chloromercuribenzene sulfonate (PCMBS).
  • 2.2. The values of Pd were around 6.3 × 10−3 cm/sec at 15°C, 7.0 × 10−3cm/sec at 20°C, 8.0 × 10−3 cm/sec at 25°C, 9.1 × 10−3 cm/sec at 30°C and10.7 × 10−3 cm/sec at 37°C.
  • 3.3. Systematic studies on the effects of PCMBS on water diffusion indicated that the maximal inhibition was reached in 15 min at 37°C with 0.5 mM PCMBS.
  • 4.4. The values of maximal inhibition were around 71–74% at all temperatures.
  • 5.5. The basal permeability to water was estimated as 1.6 × 10−3cm/sec at 15°C, 2.0 × 10−3cm/sec at 20°C, 2.4 × 10−3cm/sec at 25°C, 2.6 × 10−3cm/sec at 30°C, and 3.1× 10−3 cm/secat 37°C.
  • 6.6. The activation energy of water diffusion was around 18 kJ/mol and increased to 27 kcal/mol after incubation with PCMBS in conditions of maximal inhibition of water diffusion.
  • 7.7. The membrane polypeptide electrophoretic pattern of rabbit RBCs has been compared with its human counterpart.
  • 8.8. The rabbit membrane contained a higher amount of spectrin (bands 1 and 2), while the band 6 (glyceraldehyde-3-phosphate dehydrogenase) was markedly less intense.
  • 9.9. Considerable differences in the electrophoretic patterns of the two sources of RBC membranes appeared in the bands migrating in the band 4.5 region and in front of band 7, where some polypeptides were apparent in higher amounts in the rabbit RBC membrane.
  相似文献   

19.
With the aid of measurements of the fluorescence yield, the efficiency of the various deexcitation mechanisms of an exciton in the light-harvesting system has been determined. For this purpose, the fluorescence of dark-adapted as well as of 3-(3,4-dichlorophenyl)-1,1-dimethylurea (DCMU)-treated and preilluminated leaves of Zea mays L. was excited by single ultrashort laser pulses of different energies. The experimental results have served for the fitting of solutions of rate equations, which describe the deexcitation by linear relaxation processes like fluorescence and radiationless transitions, by annihiation of excitons, and by traps both in the ground state and in an excited state. We have obtained the following results: a ratio of antenna chlorophyll molecules to Photosystem II traps of 600:1, an annihilation constant γ = 2·10?8 cm3·s?1, a mean trapping time of t?=0.5 ns, a trapping probability for traps in the ground state of 2·10?8 cm3·s?1, and 6·10?9 cm3·s?1 for traps in an excited state.  相似文献   

20.
The calcium efflux from multi-layered vesicles (liposomes) of different lipid composition has been studied. Liposomes composed of lipids extracted from cattle retinas are compared with liposomes which consist of phosphatidylcholine or a 1 : 1 phosphatidylcholine/phosphatidylserine mixture. The percentages of 45Ca capture by these three types of liposomes are 10, 1 and 4% respectively.The efflux rates are 2.5 · 10?6, 2 · 10?6 and 4 · 10?5 s?1 respectively. The semilogarithmic efflux curves for phosphatidylcholine and phosphatidylcholine/phosphatidylserine liposomes are linear with time, but those for the retinal lipid liposomes are discontinuous. The activation energy for the calcium efflux from the latter liposomes is about 10.5 kcal/mol, both before and after the discontinuity.The ionophores X537A and A23187 enhance the calcium leakage from retinal lipid liposomes, the latter ionophore being much more effective than the former. At high concentrations both ionophores seem to transport calcium as a 1 : 2Ca · ionophore complex. At low ionophore concentrations, however, X537A appears to transport calcium as a 1 : 1 complex, but A23187 as a 2 : 1 complex.  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