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 31 毫秒
1.
Background and Aims Polyphenol oxidases (PPOs) catalyse the oxidation of monophenols and/or o-diphenols to highly reactive o-quinones, which in turn interact with oxygen and proteins to form reactive oxygen species (ROS) and typical brown-pigmented complexes. Hence PPOs can affect local levels of oxygen and ROS. Although the currently known substrates are located in the vacuole, the enzyme is targeted to the thylakoid lumen, suggesting a role for PPOs in photosynthesis. The current study was designed to investigate the potential involvement of PPOs in the photosynthetic response to oxidative stress.Methods Photosynthesis (A, Fv/Fm, ΦPSII, qN, qP, NPQ) was measured in leaves of a wild-type and a low-PPO mutant of red clover (Trifolium pratense ‘Milvus’) under control conditions and under a stress treatment designed to induce photooxidative stress: cold/high light (2 °C/580 µmol m2 s–1) or 0–10 µm methyl viologen. Foliar protein content and oxidation state were also determined.Key Results Photosynthetic performance, and chlorophyll and protein content during 4 d of cold/high light stress and 3 d of subsequent recovery under control growth conditions showed similar susceptibility to stress in both lines. However, more extensive oxidative damage to protein in mutants than wild-types was observed after treatment of attached leaves with methyl viologen. In addition, PPO activity could be associated with an increased capacity to dissipate excess energy, but only at relatively low methyl viologen doses.Conclusions The presence of PPO activity in leaves did not correspond to a direct role for the enzyme in the regulation or protection of photosynthesis under cold stress. However, an indication that PPO could be involved in cellular protection against low-level oxidative stress requires further investigation.  相似文献   

2.
Propagated action potentials can be obtained in canine cardiac Purkinje fibers exposed to Na-free solutions containing no inorganic cation other than Ca and K. Essentially similar action potentials are obtained if Na is replaced by tetraethylammonium (TEA), tetramethylammonium (TMA), or choline. In a solution containing 128 mM TEA and 16.2 mM Ca the characteristics of these electrical responses were: maximum diastolic potential, -59 ± 3.3 mV; overshoot, 20 ± 6.8 mV; maximum upstroke velocity, 3.7 ± 2.3 V/s; conduction velocity, 0.1 m/s; and action potential duration, 360 ± 45 ms. The magnitude of the overshoot varied with log Cao with a slope of about 30 mV/10-fold concentration change. The upstroke velocity was an approximately linear function of Cao. The active response was greatly diminished or abolished by Mn and D-600 but was unaffected by tetrodotoxin. These Ca-dependent responses appeared in a region of transmembrane potential (about -50 mV) at which the rapid Na-dependent upstroke is abolished even when Na is present.  相似文献   

3.
1. Enzyme systems from Cucurbita pepo have been shown to catalyse the reduction of nitrite and hydroxylamine to ammonia in yields about 90–100%. 2. Reduced benzyl viologen serves as an efficient electron donor for both systems. Activity of the nitrite-reductase system is directly related to degree of dye reduction when expressed in terms of the function for oxidation–reduction potentials, but appears to decrease to negligible activity below about 9% dye reduction. 3. NADH and NADPH alone produce negligible nitrite loss, but NADPH can be linked to an endogenous diaphorase system to reduce nitrite to ammonia in the presence of catalytic amounts of benzyl viologen. 4. The NADH– or NADPH–nitrate-reductase system that is also present can accept electrons from reduced benzyl viologen, but shows relationships opposite to that for the nitrite-reductase system with regard to effect of degree of dye reduction on activity. The product of nitrate reduction may be nitrite alone, or nitrite and ammonia, or ammonia alone, according only to the degree of dye reduction. 5. The relative activities of nitrite-reductase and hydroxylamine-reductase systems show different relationships with degree of dye reduction and may become reversed in magnitude when effects of degree of dye reduction are tested over a suitable range. 6. Nitrite severely inhibits the rate of reduction of hydroxylamine without affecting the yield of ammonia as a percentage of total substrate loss, but hydroxylamine has a negligible effect on the activity of the nitrite-reductase system. 7. The apparent Km for nitrite (1 μm) is substantially less than that for hydroxylamine, for which variable values between 0·05 and 0·9mm (mean 0·51 mm) have been observed. 8. The apparent Km values for reduced benzyl viologen differ for the nitrite-reductase and hydroxylamine-reductase systems: 60 and 7·5 μm respectively. 9. It is concluded that free hydroxylamine may not be an intermediate in the reduction of nitrite to ammonia by plants, and a possible mechanism for reduction of both compounds by the same enzyme system is discussed in the light of current ideas relating to other organisms.  相似文献   

4.
Mannan RM  Bose S 《Plant physiology》1986,80(1):264-268
When Triticum vulgare cv HD 2189 seedlings were grown in the presence of 125 micromolar BASF 13.338 (4-chloro-5-dimethylamino-2-phenyl-3(2H)pyridazinone), the rate of electron transport (H2O → methyl viologen) in chloroplast thylakoids isolated from the treated seedlings was higher (by 50%) as compared to the control at assay temperatures above 30°C. Below 30°C, however, the rate with the treated seedlings was lower than the control rate. The temperature dependence of the rate of photosystem I electron transport (2-6-dichlorophenol indophenol-reduced → methyl viologen) in the treated system was similar to that in the control. At high temperatures (>30°C), with diphenyl carabazide as electron donor, the rates of electron transfer (diphenyl carbazide → methyl viologen) were similar in the treated and in the control thylakoids. Direct addition of BASF 13.338 to the assay mixture for the measurement of rate of electron transport (H2O → methyl viologen) in the thylakoids isolated from the control plants did not cause any change in the temperature dependence of photosynthetic electron transport. These results suggested that the donor side of photosystem II became tolerant to heat in the treated plants. Chlorophyll a fluorescence emission was monitored continuously in the leaves of control and BASF 13.338 treated wheat seedlings during continuous increase in temperature (1°C per minute). The fluorescence-temperature profile showed a decrease in the fluorescence yield above 55°C; this decrease was biphasic in the control and monophasic in the treated plants.  相似文献   

5.
The cells of cat right ventricular papillary muscles were depleted of K and caused to accumulate Na and water by preincubation at 2–3°C. The time courses of changes in cellular ion content and volume and of the resting membrane potential (Vm) were then followed after abrupt rewarming to 27–28°C. At physiological external K concentration ([K]o = 5.32 mM) recovery of cellular ion and water contents was complete within 30 minutes, the maximal observable rates of K uptake and Na extrusion (Δmmol cell ion/(kg dry weight) (min.)) being 3.4 and 3.6, respectively. The recovery rate was markedly slowed at [K]o = 1.0 mM. Rewarming caused Vm measured in cells at the muscle surface to recover within from <1 to 9 minutes, but only slight restoration of cellular ion contents (measured in whole muscles) had occurred after 10 minutes. Studies of recovery in NaCl-free sucrose Ringer''s solution made it possible to separate the ouabain-insensitive outward diffusion of Na as a salt from a simultaneous ouabain-sensitive Na extrusion which is associated with a net cellular K uptake. A hypothesis consistent with these observations is that rewarming may activate a ouabain-sensitive "electrogenic" mechanism, most probably the net active transport of Na out of the cell, from which net K uptake may then follow passively.  相似文献   

6.
Suspensions of the yeast Saccharomyces cerevisiae gave reproducible rates of O2 uptake over a period of 6 months. The relation of rate of consumption of O2 to temperature was tested over a wide range of temperatures, and the constant in the formulation of the relationship is found to be reproducible. The values of this constant (µ) have been obtained for five separate series of experiments by three methods of estimation. The variability of µ has the following magnitudes: the average deviation of a single determination expressed as per cent of the mean is ±2 per cent in the range 30–15°, and ±0.8 per cent in the range 15–3°C. This constancy of metabolic activity measured as a function of temperature can then be utilized for more precise investigations of processes controlling the velocity of oxidations of substrates, and of respiratory systems controlled by intracellular respiratory pigments. The data plotted according to the Arrhemus equation give average values of the constant µ as follows: for the range 35–30°, µ = 8,290; 30–15°, µ = 12,440 ±290; 15–3°, µ = 19,530 ±154. The critical temperatures are at 29.0° and 15.7°C. A close similarity exists between these temperature characteristics (µ) and values in the series usually obtained for respiratory activities in other organisms. This fact supports the view that a common system of processes controls the velocities of physiological activities in yeast and in other organisms.  相似文献   

7.
1. Oxidation-reduction potential methods have been applied to a study of the blue-purple pigment present in solution in the blood and in the tissue cells of the nudibranch Chromodoris zebra. 2. The blue-purple pigment and its yellow reduction product form a reversible system whose Eo'' = x0.102 volts at pH 7.0 and whose valence change from oxidant to reductant appears to be one. 3. The system is unlike oxyhemoglobin-hemoglobin in the mode of oxygen transfer. Its rôle as a possible respiratory material is discussed.  相似文献   

8.
The electrophysiological properties of cytoplasm-rich fragments (single membrane samples) prepared from internodal cells of Chara corallina were explored in conjunction with K+-sensitive microelectrode and current-voltage (I-V) measurements. This system eliminated the problem of the inaccessible cytoplasmic layer, while preserving many of the electrical characteristics of the intact cells. In 0.1 millimolar external K concentration (Ko+), the resting conductance (membrane conductance Gm, 0.85 ± 0.25 Siemens per square meter (±standard error)) of the single membrane samples, was dominated by the proton pump, as suggested by the response of the near-linear I-V characteristic to changes in external pH. Initial cytoplasmic K+ activities (aK+), judged most reliable, gave values of 117 ± 67 millimolar; stable aK+ values were 77 ± 31 millimolar. Equilibrium potentials for K+ (Nernst equilibrium potential) (EK) calculated, using either of these data sets, were near the mean membrane potential (Vm). On a cell-to-cell basis, however, EK was generally negative of the Vm, despite an electrogenic contribution from the Chara proton pump. When Ko+ was increased to 1.0 millimolar or above, Gm rose (by 8- to 10-fold in 10 millimolar Ko+), the steady state I-V characteristics showed a region of negative slope conductance, and Vm followed EK. These results confirm previous studies which implicated a Ko+-induced and voltage-dependent permeability to K+ at the Chara plasma membrane. They provide an explanation for transitions between apparent Ko+-insensitive and Ko+-sensitive (`K+ electrode') behavior displayed by the membrane potential, as recorded in many algae and higher plant cells.  相似文献   

9.
In the labellar chemosensory hairs of the blowfly, Phormia regina Meigen, stationary amplitudes of the slow potentials induced by salt and sugar stimulations were decreased to 50–80% at 12°C of the values measured at 28°C. The amplitudes induced by water did not show any dependence on temperature change. The maximum rate of rise of the receptor potentials was strongly increased with rising temperature. The value of Kb, the apparent Michaelis constant, was less by a factor of six at 28°C compared to the 12°C value for the sugar receptor.  相似文献   

10.
The effect of temperature on the potential and current thresholds of the squid giant axon membrane was measured with gross external electrodes. A central segment of the axon, 0.8 mm long and in sea water, was isolated by flowing low conductance, isoosmotic sucrose solution on each side; both ends were depolarized in isoosmotic KCl. Measured biphasic square wave currents at five cycles per second were applied between one end of the nerve and the membrane of the central segment. The membrane potential was recorded between the central sea water and the other depolarized end. The recorded potentials are developed only across the membrane impedance. Threshold current values ranged from 3.2 µa at 267deg;C to 1 µa at 7.5°C. Threshold potential values ranged from 50 mv at 26°C to 6 mv at 7.5°C. The mean Q10 of threshold current was 2.3 (SD = 0.2), while the Q10 for threshold potentials was 2.0 (SD = 0.1).  相似文献   

11.
The total CO2 produced by aseptic Drosophila cultures during the entire duration of life has been determined at 15°, 26°, and 30°C. in the dark and at 22–26°C. in the light. The total amount of CO2 produced is not constant but is greater at 15° than at 26° or 30°, and is much greater in the light than in the dark. The total duration of life, therefore, is not determined by the time required to produce a limiting amount of CO2.  相似文献   

12.
Permeation of cystic fibrosis transmembrane conductance regulator (CFTR) Cl channels by halide ions was studied in stably transfected Chinese hamster ovary cells by using the patch clamp technique. In cell-attached patches with a high Cl pipette solution, the CFTR channel displayed outwardly rectifying currents and had a conductance near the membrane potential of 6.0 pS at 22°C or 8.7 pS at 37°C. The current–voltage relationship became linear when patches were excised into symmetrical, N-tris(hydroxymethyl)methyl-2-aminomethane sulfonate (TES)-buffered solutions. Under these conditions, conductance increased from 7.0 pS at 22°C to 10.9 pS at 37°C. The conductance at 22°C was ∼1.0 pS higher when TES and HEPES were omitted from the solution, suggesting weak, voltage-independent block by pH buffers. The relationship between conductance and Cl activity was hyperbolic and well fitted by a Michaelis-Menten–type function having a K m of ∼38 mM and maximum conductance of 10 pS at 22°C. Dilution potentials measured with NaCl gradients indicated high anion selectivity (PNa/PCl = 0.003–0.028). Biionic reversal potentials measured immediately after exposure of the cytoplasmic side to various test anions indicated PI (1.8) > PBr (1.3) > PCl (1.0) > PF (0.17), consistent with a “weak field strength” selectivity site. The same sequence was obtained for external halides, although inward F flow was not observed. Iodide currents were protocol dependent and became blocked after 1–2 min. This coincided with a large shift in the (extrapolated) reversal potential to values indicating a greatly reduced I/Cl permeability ratio (PI/PCl < 0.4). The switch to low I permeability was enhanced at potentials that favored Cl entry into the pore and was not observed in the R347D mutant, which is thought to lack an anion binding site involved in multi-ion pore behavior. Interactions between Cl and I ions may influence I permeation and be responsible for the wide range of PI/PCl ratios that have been reported for the CFTR channel. The low PI/PCl ratio usually reported for CFTR only occurred after entry into an altered permeability state and thus may not be comparable with permeability ratios for other anions, which are obtained in the absence of iodide. We propose that CFTR displays a “weak field strength” anion selectivity sequence.  相似文献   

13.
In previous papers of this series the temperature-dependent Raman spectra of poly(dA)·poly(dT) and poly(dA–dT)·poly(dA–dT) were used to characterize structurally the melting and premelting transitions in DNAs containing consecutive A·T and alternating A·T/T·A base pairs. Here, we describe procedures for obtaining thermodynamic parameters from the Raman data. The method exploits base-specific and backbone-specific Raman markers to determine separate thermodynamic contributions of A, T and deoxyribosyl-phosphate moieties to premelting and melting transitions. Key findings include the following: (i) Both poly(dA)·poly(dT) and poly(dA–dT)· poly(dA–dT) exhibit robust premelting transitions, due predominantly to backbone conformational changes. (ii) The significant van’t Hoff premelting enthalpies of poly(dA)·poly(dT) [ΔHvHpm = 18.0 ± 1.6 kcal·mol–1 (kilocalories per mole cooperative unit)] and poly(dA–dT)·poly(dA–dT) (ΔHvHpm = 13.4 ± 2.5 kcal·mol–1) differ by an amount (~4.6 kcal·mol–1) estimated as the contribution from three-centered inter-base hydrogen bonding in (dA)n·(dT)n tracts. (iii) The overall stacking free energy of poly(dA)· poly(dT) [–6.88 kcal·molbp–1 (kilocalories per mole base pair)] is greater than that of poly(dA–dT)· poly(dA–dT) (–6.31 kcal·molbp–1). (iv) The difference between stacking free energies of A and T is significant in poly(dA)·poly(dT) (ΔΔGst = 0.8 ± 0.3 kcal· molbp–1), but marginal in poly(dA–dT)·poly(dA–dT) (ΔΔGst = 0.3 ± 0.3 kcal·molbp–1). (v) In poly(dA)· poly(dT), the van’t Hoff parameters for melting of A (ΔHvHA = 407 ± 23 kcal·mol–1, ΔSvHA = 1166 ± 67 cal·°K–1·mol–1, ΔGvH(25°C)A = 60.0 ± 3.2 kcal·mol–1) are clearly distinguished from those of T (ΔHvHT = 185 ± 38 kcal·mol–1, ΔSvHT = 516 ± 109 cal·°K–1·mol–1, ΔGvH(25°C)T = 27.1 ± 5.5 kcal·mol–1). (vi) Similar relative differences are observed in poly(dA–dT)· poly(dA–dT) (ΔHvHA = 333 ± 54 kcal·mol–1, ΔSvHA = 961 ± 157 cal·°K–1·mol–1, ΔGvH(25°C)A = 45.0 ± 7.6 kcal· mol–1; ΔHvHT = 213 ± 30 kcal·mol–1, ΔSvHT = 617 ± 86 cal·°K–1·mol–1, ΔGvH(25°C)T = 29.3 ± 4.9 kcal·mol–1). The methodology employed here distinguishes thermodynamic contributions of base stacking, base pairing and backbone conformational ordering in the molecular mechanism of double-helical B DNA formation.  相似文献   

14.
Efflux and Influx of Erythrocyte Water   总被引:1,自引:1,他引:0       下载免费PDF全文
Rabbit erythrocytes were washed in buffered NaCl solutions isotonic with rabbit serum (Δt -0.558°C.) and suspended in buffered NaCl solutions of tonicity equidistant from intracellular tonicity (Δt = -0.558°C. ± 0.112°C.) of varying pH and incubated at varying temperatures. After incubation, the freezing point depression (Δt) was measured on the supernatant. Change in the Δt measured change in the water content of the extracellular solutions—water being withdrawn by erythrocytes (WI) from the hypotonic solutions and added (WE) to the hypertonic solutions. WE was always less than WI and was inversely proportional to the pH in the range 6.5–8.0. WE was significantly increased by lowering the temperature of the cell suspension to 4°C. WI was increased by raising or lowering the pH or raising the temperature of the cell suspension. WE x WIk. WE and WI were affected differently by changes in pH and temperature. It was concluded that WE and WE were probably under different physicochemical control.  相似文献   

15.
Current-voltage curves for DIDS-insensitive Cl conductance have been determined in human red blood cells from five donors. Currents were estimated from the rate of cell shrinkage using flow cytometry and differential laser light scattering. Membrane potentials were estimated from the extracellular pH of unbuffered suspensions using the proton ionophore FCCP. The width of the Gaussian distribution of cell volumes remained invariant during cell shrinkage, indicating a homogeneous Cl conductance among the cells. After pretreatment for 30 min with DIDS, net effluxes of K+ and Cl were induced by valinomycin and were measured in the continued presence of DIDS; inhibition was maximal at ∼65% above 1 μM DIDS at both 25°C and 37°C. The nonlinear current-voltage curves for DIDS-insensitive net Cl effluxes, induced by valinomycin or gramicidin at varied [K+]o, were compared with predictions based on (1) the theory of electrodiffusion, (2) a single barrier model, (3) single occupancy, multiple barrier models, and (4) a voltage-gated mechanism. Electrodiffusion precisely describes the relationship between the measured transmembrane voltage and [K+]o. Under our experimental conditions (pH 7.5, 23°C, 1–3 μM valinomycin or 60 ng/ml gramicidin, 1.2% hematocrit), the constant field permeability ratio PK/PCl is 74 ± 9 with 10 μM DIDS, corresponding to 73% inhibition of PCl. Fitting the constant field current-voltage equation to the measured Cl currents yields P Cl = 0.13 h−1 with DIDS, compared to 0.49 h−1 without DIDS, in good agreement with most previous studies. The inward rectifying DIDS-insensitive Cl current, however, is inconsistent with electrodiffusion and with certain single-occupancy multiple barrier models. The data are well described either by a single barrier located near the center of the transmembrane electric field, or, alternatively, by a voltage-gated channel mechanism according to which the maximal conductance is 0.055 ± 0.005 S/g Hb, half the channels are open at −27 ± 2 mV, and the equivalent gating charge is −1.2 ± 0.3.  相似文献   

16.
1. A group of normal and congenitally goitrous Merino sheep were investigated to identify the metabolic defect present in the abnormal animals. 2. Protein-bound iodine concentrations of serum from goitrous animals (average 5·7μg./100ml.) were higher than normal (average 4·2μg./100ml.; P 0·001), but the hormonal iodine measured as butanol-extractable 131I was low in the serum of goitrous (average 40·3% of protein-bound 131I) compared with that of normal (84·2%; P 0·02) sheep. The non-hormonal iodine of the serum of goitrous sheep appeared to include iodotyrosines and iodinated protein. 3. Starch-gel-electrophoretic separations of sera from normal and goitrous sheep after 131I injection (100–500μc) showed no qualitative differences in the radioactivity of protein components. No significant differences in thyroxine-binding in vitro by serum proteins of normal and goitrous sheep were observed. 4. The clearance rates of 131I-labelled iodotyrosines (t½ 1·2–2·9hr.) and iodothyronines (t½ 33·5–47·4hr.) were similar in normal and goitrous sheep. 5. The concentration of circulating thyroid-stimulating hormone was significantly higher (P<0·01 in three sheep, P<0·05 in one sheep) in goitrous sheep. 6. The congenital goitre appears to be due to compensatory hypertrophy of the gland resulting from an inability to synthesize an adequate supply of thyroid hormone.  相似文献   

17.
Monocular threshold stimulus intensities (ΔIo, photons) were measured along the 0–180° meridian of human retinae for three observers. The test image was small (= 0.08°) and of short duration (= 0.20 second). ΔIo was found to decrease as the angular distance from the fovea was increased. Actual counts of the number of retinal elements per mm.2 along the 0–180° meridian (Østerberg) were compared with the obtained results. No direct correlation was found to exist between visual sensitivity and the number of retinal elements. Binocular threshold stimuli were also measured along the same meridian. The form of the function relating binocular visual sensitivity and retinal position was discovered to be essentially similar to that for monocular sensitivity, but is more symmetrical about the center of the fovea. The magnitude of the binocular measurement is in each case smaller than that of the monocular threshold stimulus intensity for the more sensitive eye. The ratio is statistically equal to 1.4 (a fact which suggests Piper''s rule). These results are shown to be consistent with the hypothesis that the process critical for the eventuation of the threshold response is localized in the central nervous system. They are not consistent with the view that the quantitative properties of visual data are directly determined by properties of the peripheral retina.  相似文献   

18.
The effect of electric fields on dry oriented multibilayers of dimyristoylphosphatidylcholine (DMPC) was investigated by transmission Fourier transform infrared electric field modulated excitation (E-ME) spectroscopy. A periodic rectangular electric potential (0–150 V, 1.25 Hz, 28.4°C ± 0.2°C) was applied across the sample. To discriminate electric field-induced effects from possible temperature-induced effects resulting from a current flow (<1 pA) across the sample, corresponding temperature-modulated excitation (T-ME) measurements within the temperature uncertainty limits of ±0.2°C at 28.4°C were performed. T-ME induced reversible gauche defects in the hydrocarbon chains, whereas E-ME resulted in reversible compression of dry DMPC bilayers. Periodic variation of the tilt angle of the hydrocarbon chains is suggested. The degree of absorbance modulation in the CH-stretching region was found to be in the order of 1:700, corresponding to a variation of the bilayer thickness of Δz = 0.0054 nm. Using a series connection of capacitors as equivalent circuit of the cell resulted in E = (1.2 ± 0.7) × 107 V/m for the electric field in DMPC. Young's elasticity modulus of DMPC could be calculated to be E = 2.2 × 106 Pa ± 1.8 × 106 Pa, which is in good agreement with published data obtained by electric field-dependent capacitance measurements.  相似文献   

19.
1. Polyphenol oxidase (EC 1. 10. 3.–) from the shoots of the tea plant was purified about 5000-fold on a dry-weight basis. 2. At an intermediate stage of purification four soluble yellow fractions were obtained. They are believed to represent complexes of a basic enzyme protein with acidic phenolic oxidation products and nucleic acids. After removal of the complex-forming materials the fractions were blue and similar to each other. About 40% of the activity could not be extracted from the acetone-dried powder. 3. Each of the four blue fractions was resolved further into two species, A and B. The following results refer to species A. 4. The enzyme showed absorption maxima at 279mμ (E1%1cm., 13·5) and 611mμ (E1%1cm., 0·84) with a shoulder at 330mμ. The enzyme was bleached by substrate under anaerobic conditions and the colour was restored by oxygen. 5. The molecular weight measured by sedimentation and diffusion was 144000±16000. The copper content was 0·32% (w/w). 6. Kinetic constants are given for a number of substrates and inhibitors, including the natural substrates of the tea leaf. The specific activity towards pyrogallol was 373 units/mg. at 30°. 7. The best substrates were o-dihydric phenols. Quinol and p-phenylenediamine were slowly oxidized. Monohydric phenols and ascorbic acid were not oxidized. 8. The kinetics of oxidation of most substrates are consistent with a mechanism in which oxidized and reduced forms of the enzyme form binary complexes with phenol and oxygen respectively. A modified mechanism is postulated for the oxidation of chlorogenic acid. 9. The relation of the results to the mechanism of tea fermentation is discussed.  相似文献   

20.
Thermostability of the photosynthetic apparatus of abscisic acid (ABA)-treated seedlings of barley (Hordeum vulgare) was studied by light-scattering and by fluorescence measurements of isolated chloroplasts. ABA treatment markedly decreased heat damage of the chloroplast ultrastructure; an exogenous ABA concentration of 10−5 molar was most effective. Heat-induced increase of the 77 kilodalton fluorescence ratio F740/F685 was also smaller at this ABA concentration. The heat-induced increase of the initial chlorophyll fluorescence level (Fo) was virtually eliminated in ABA-treated (10−5 molar) chloroplasts up to 45°C and slightly increased at 50°C, relative to control chloroplasts where Fo increased even at 35°C and reached its maximal value at 45°C. In control chloroplasts, Fo increased with a 5-minute pretreatment temperature, an effect observed as low as 35°C. Fo was maximal at 45°C. In contrast, chloroplasts treated with 10−5 molar ABA did not exhibit a heat-induced increase in Fo until 50°C.  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