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.
Field bean (Dolichos lablab) contains a single isoform of PPO (polyphenol oxidase)--a type III copper protein that catalyses the o-hydroxylation of monophenols and oxidation of o-diphenols using molecular oxygen--and is a homotetramer with a molecular mass of 120 kDa. The enzyme is activated manyfold either in the presence of the anionic detergent SDS below its critical micellar concentration or on exposure to acid-pH. The enhancement of kcat upon activation is accompanied by a marked shift in the pH optimum for the oxidation of t-butyl catechol from 4.5 to 6.0, an increased sensitivity to tropolone, altered susceptibility to proteolytic degradation and decreased thermostability. The Stokes radius of the native enzyme is found to increase from 49.1+/-2 to 75.9+/-0.6 A (1 A=0.1 nm). The activation by SDS and acid-pH results in a localized conformational change that is anchored around the catalytic site of PPO that alters the microenvironment of an essential glutamic residue. Chemical modification of field bean and sweet potato PPO with 1-ethyl-3-(3-dimethylaminopropyl)carbodi-imide followed by kinetic analysis leads to the conclusion that both the enzymes possess a core carboxylate essential to activity. This enhanced catalytic efficiency of PPO, considered as an inducible defence oxidative enzyme, is vital to the physiological defence strategy adapted by plants to insect herbivory and pathogen attack.  相似文献   

2.
Polyphenol oxidase (PPO) was extracted from beet root, in both soluble and membrane fractions, and in both cases the enzyme was in a latent state. PPO from the membrane fraction showed no diphenolase activity unless it was activated by trypsin or sodium dodecyl sulfate (SDS). The kinetics of the activation process of latent PPO by trypsin was studied and the specific rate constant of active PPO formation, k 3 , showed a value of 0.03 s(-1). The protease-activated form showed a pH optimum (6.5) and kinetic properties identical to those of the SDS-activated enzyme. Evidence is provided for the existence of a common peptide responsible for the regulation of the activity of the enzyme by both proteolysis and SDS detergent. Formation of the active proteolyzate was followed by spectroscopic measurements, Western blotting and partially denaturing SDS-PAGE.  相似文献   

3.
The monophenolase activity of Terfezia claveryi tyrosinase (EC 1.14.18.1) is described for the first time. This enzyme is fully latent and can only be detected if SDS is present in the reaction medium. Monophenolase activity was localized within the ascocarp using histochemical techniques. A detailed kinetic study of the parameters affecting this activity has been carried out. Both the characteristic lag period and the steady-state rate are affected by pH and the enzyme and substrate concentrations. The presence of catalytic concentrations of o -diphenols affected the lag period but not the steady-state rate. By increasing the concentration of o- diphenols, it was possible to evaluate the enzyme activation constant, Kact, which showed a value of 7.2 μ M . The experimental results are compatible with the mechanism previously described for tyrosinases from other sources.  相似文献   

4.
The 20S proteasome of eukaryotic cells has at least three distinct peptidase activities (trypsin-like, chymotrypsin-like and peptidylglutamylpeptide (PGP) hydrolase activities). These peptidases are latent and require appropriate activators. SDS has been widely used as an activator of these peptidases, but the mechanism of its activation remains unresolved. In this study, we investigated the kinetics of the SDS-activated hydrolysis of the above three types of peptidase of the 20S proteasome purified from Xenopus oocytes. When the reaction was started by simultaneous adding both SDS and substrate, maximal rates of hydrolysis were reached after appreciable lag phases with the trypsin-type substrate [t-butyloxycarbonylLeu-Arg-Arg-4-methylcoumaryl-7-amide (Boc-LRR-MCA)], but no such lag phases were observed with the chymotrypsin-type and PGP hydrolase-type substrates [succinyl-Leu-Leu-Val-Tyr-4-methylcoumaryl-7-amide (Suc-LLVY-MCA), and benzyloxycarbonyl-Leu-Leu-Glu-2-naphthylamide (Cbz-LLE-2NA), respectively]. Similarly, changes in the hydrolysis rate to a reduced level upon dilution of SDS occurred after an appreciable lag phase again in the trypsin-like peptidase, but not in the other types. The lag phase characteristic of the trypsin-like peptidase was dependent on the substrate concentration. Thus, the lag phase was less discernible at very low concentrations of the substrate (e.g. at concentrations in the order of 1/100 of the Km value), but became more conspicuous with the increases in the substrate concentration. This lag phase also vanished upon preincubation of the activator (SDS) for a short period of 5 sec. These results suggest that the formation of the enzyme-substrate complex in the trypsin-like reaction induces a conformational change in the enzyme which makes the SDS activator site(s) in an occluded form, reducing the rates of SDS binding and dissociation.  相似文献   

5.
The effect of sodium dodecyl sulfate (SDS) on purified tobacco leaf PPO (PPO II) was investigated at various pHs and temperatures. SDS increased the activity of PPO II due to the formation of SDS-PPO II complex, leading to conformational changes, thus making access to active center easier. The relationship between the activity and the molar ratio of SDS-PPO II to PPO II showed that the critical point reached a plateau of activity at the molar ratio of about 1.2. The pH had a significant effect on interaction between SDS and PPO II, as compared to PPO II. The optimum catalytic temperature of the complex rose by 10 degrees C, suggesting that stabilization of the structure had been improved by the formation of complex.  相似文献   

6.
Appearance of a lag period dependent on pH in the expression of the catecholase activity of a polyphenoloxidase extracted in a latent state from Airen grape (Vitis vinifera L.) berries, is revealed, suggesting the hysteretic nature of the enzyme. The lag time was independent of enzyme concentration, indicating that slow pH-induced conformational changes in the protein must occur during assay. Results obtained by varying substrate concentration show that the system presents hyperbolic or cooperative kinetics depending on the pH of the assay.  相似文献   

7.
Highly purified pyruvic decarboxylase (EC 4.1.1.1) from wheat germ catalyses the decarboxylation of hydroxypyruvate. A kinetic analysis of the activity of the enzyme with pyruvate and hydroxypyruvate as substrates suggests that a single enzyme is involved. The kinetics of decarboxylation are autocatalytic. The time lag before maximum activity is reached is affected by the concentration of hydroxypyruvate and the pH. The question whether or not hydroxypyruvate is a natural substrate for the enzyme remains unresolved, but it may be significant that at physiological pH (ca 7.5) the enzyme shows optimum activity with hydroxypyruvate, but negligible activity with pyruvate.  相似文献   

8.
The kinetics of the NAD+-dependent oxidation of aldehydes, catalysed by aldehyde dehydrogenase purified from sheep liver mitochondria, were studied in detail. Lag phases were observed in the assays, the length of which were dependent on the enzyme concentration. The measured rates after the lag phase was over were directly proportional to the enzyme concentration. If enzyme was preincubated with NAD+, the lag phase was eliminated. Double-reciprocal plots with aldehyde as the variable substrate were non-linear, showing marked substrate activation. With NAD+ as the variable substrate, double-reciprocal plots were linear, and apparently parallel. Double-reciprocal plots with enzyme modified with disulfiram (tetraethylthiuram disulphide) or iodoacetamide, such that at pH 8.0 the activity was decreased to 50% of the control value, showed no substrate activation, and the plots were linear. At pH 7.0, the kinetic parameters Vmax. and Km NAD+- for the oxidation of acetaldehyde and butyraldehyde by the native enzyme are almost identical. Formaldehyde and propionaldehyde show the same apparent maximum rate. Aldehyde dehydrogenase is able to catalyse the hydrolysis of p-nitrophenyl esters. This esterase activity was stimulated by both NAD+ and NADH, the maximum rate for the NAD+ stimulated esterase reaction being roughly equal to the maximum rate for the oxidation of aldehydes. The mechanistic implications of the above behaviour are discussed.  相似文献   

9.
A lag is observed before the steady state during pyruvate reduction catalysed by lactate dehydrogenase from Streptococcus lactis. The lag is abolished by preincubation of enzyme with the activator fructose 1,6-bisphosphate before mixing with the substrates. The rate constants for the lag phase showed a linear dependence on fructose-1,6-bisphosphate concentration, with a second-order rate constant of 2.0 X 10(4) M-1 s-1, but were independent of enzyme concentration. Binding of fructose 1,6-bisphosphate produces a decrease in the protein fluorescence of the enzyme. The second-order rate constant for the fluorescence change is twice that for the lag in pyruvate reduction. The results suggest that binding of fructose 1,6-bisphosphate induces a conformational change in the enzyme, producing a form with reduced protein fluorescence and increased activity towards pyruvate reduction.  相似文献   

10.
The effects of detergents, trypsin and fatty acids on structural and functional properties of a pure loquat fruit latent polyphenol oxidase have been studied in relation to its regulation. Anionic detergents activated PPO at pH 6.0 below critical micelle concentration (cmc), but inhibited at pH 4.5 well above cmc. This behavior is due to a detergent-induced pH profile alkaline shift, accompanied by changes of intrinsic fluorescence of the protein. Gel filtration experiments demonstrate the formation of PPO-SDS mixed micelles. Partial PPO proteolysis suggest that latent PPO losses an SDS micelle-interacting region but conserves an SDS monomer-interacting site. Unsaturated fatty acids inhibit PPO at pH 4.5, the strongest being linolenic acid while the weakest was gamma-linolenic acid for both, the native and the trypsin-treated PPO. Down-regulation of PPO activity by anionic amphiphiles is discussed based on both, the pH profile shift induced upon anionic amphiphile binding and the PPO interaction with negatively charged membranes.  相似文献   

11.
A wounding-induced PPO from cowpea (Vigna unguiculata) seedlings   总被引:1,自引:0,他引:1  
Polyphenol oxidases (PPO) are induced in cowpea plants by wounding. The highest activity levels were detected 48h after this stimulus in both wounded and neighbor-to-wounded unifoliates of cowpea seedlings; the increase of activity was in the order of 13 to 15-fold, respectively, in comparison to control unifoliates. Multiple molecular forms of active PPO (Mrs 58, 73 and congruent with220kDa) were detected by partially denaturing SDS-PAGE. Wounding-induced cowpea PPO were extracted and purified through (NH(4))(2)SO(4) precipitation and ion-exchange chromatography. The effects of substrate specificity, pH, thermal stability and sensitivity to various inhibitors - resorcinol, EDTA, sodium azide and tropolone - of partially purified soluble PPO were investigated. Purified wounding-induced cowpea PPO (wicPPO) showed the highest activities towards 4-methylcatechol (K(m)=9.86mM, V(max)=24.66 EU [DeltaAmin(-1)]) and catechol (K(m)=3.44mM, V(max)=6.64 EU [DeltaAmin(-1)]); no activity was observed towards l-tyrosine, under the assay conditions used. The optimum pH for wound-induced cowpea PPO was 6.0 with 4-methylcatechol as substrate. The enzyme was optimally activated by 10 mM SDS and was highly stable even after 5 min at 80 degrees C. The most effective inhibitor was tropolone, whereas addition of 10mM of resorcinol, EDTA and sodium azide were able to reduce PPO activities by 40%, 15% and 100%, respectively.  相似文献   

12.
Polyphenol oxidase (PPO) of Thymbra (Thymbra spicata L. var. spicata) was isolated by (NH4)2SO4 precipitation and dialysis. A diphenolase from Thymbra plant, active against 4-methylcatechol, catechol and pyrogallol was characterized in detail in terms of pH and temperature optima, stability, kinetic parameters and inhibition behaviour towards some general PPO inhibitors. 4-Methylcatechol was the most suitable substrate, due to the lowest Km and the biggest Vmax/Km values, followed by catechol and pyrogallol. The Thymbra PPO had maximum activity at pH 5.0, 7.0 and 8.0 with 4-methylcatechol, catechol and pyrogallol substrates, respectively. The optimum temperature of activity for Thymbra PPO was 30, 40 and 50 °C for 4-methylcatechol, catechol and pyrogallol substrates, respectively. It was found that optimum temperature and pH were substrate-dependent studied. The enzyme activity decreased due to heat denaturation of the enzyme with increasing temperature and inactivation time. Inhibition of Thymbra PPO was investigated with inhibitors such as l-cysteine and glutathione using 4-methylcatechol, catechol and pyrogallol as substrates. It was found that l-cysteine was a more effective inhibitor than glutathione owing to lower Ki. The type of inhibition depended on the origin of the PPO studied and also on the substrate used. Furthermore, the IC50 values of inhibitors sudied on PPO were determined by means of activity percentage (I) diagrams.  相似文献   

13.
Jack bean urease (urea aminohydrolase, EC 3.5.1.5) was immobilized onto modified non-porous poly(ethylene glycol dimethacrylate/2-hydroxy ethylene methacrylate), (poly(EGDMA/HEMA)), microbeads prepared by suspension copolymerization for the potential use in hemoperfusion columns, not previously reported. The conditions of immobilization; enzyme concentration, medium pH, substrate and ethylene diamine tetra acetic acid (EDTA) presence in the immobilization medium in different concentrations, enzyme loading ratio, processing time and immobilization temperature were investigated for highest apparent activity. Immobilized enzyme retained 73% of its original activity for 75 days of repeated use with a deactivation constant kd = 3.72 x 10(-3) day(-1). A canned non-linear regression program was used to estimate the intrinsic kinetic parameters of immobilized enzyme with a low value of observable Thiele modulus (phi < 0.3) and these parameters were compared with those of free urease. The best-fit kinetic parameters of a Michaelis-Menten model were estimated as Vm = 3.318 x 10(-4) micromol/s mg bound enzyme protein, Km = 15.94 mM for immobilized, and Vm = 1.074 micromol NH3/s mg enzyme protein, Km = 14.49 mM for free urease. The drastic decrease in Vm value was attributed to steric effects, conformational changes in enzyme structure or denaturation of the enzyme during immobilization. Nevertheless, the change in Km value was insignificant for the unchanged affinity of the substrate with immobilization. For higher immobilized urease activity, smaller particle size and concentrated urease with higher specific activity could be used in the immobilization process.  相似文献   

14.
Polyphenol oxidase (PPO) has been extracted from both soluble and particulate fractions of loquat fruit (Eriobotrya japonica Lindl. cv. Algerie). The soluble PPO (20% of total activity) was partially purified 3.3-fold after ammonium sulfate fractionation being in its active state. The particulate PPO fraction (80% of total activity) was purified to homogeneity in a latent form being activable by sodium dodecyl sulfate (SDS). The enzyme was purified 40.0-fold with a total yield of 15.3% after extraction by phase partitioning in Triton X-114 followed by three chromatographic steps. The molecular weight was estimated to be about 59.2 and 61.2 kDa by sodium dodecyl sulfate-polyacrylamide gel electrophoresis (SDS-PAGE) and gel filtration chromatography, respectively, indicating that latent PPO is a monomer. Latent PPO catalyzed the oxidation of chlorogenic acid (CA) at a rate 50-fold faster than that of 4-tert-butylcatechol (TBC) but the soluble active counterpart only twice. Both PPOs exhibited similar Km values for TBC but Km for CA was 5-fold higher for the latent than for the active soluble PPO. Other kinetic characteristics, including sensitivity to inhibitors, substrate specificity, thermal stability, temperature, and pH profiles, were quite different between both PPOs. These results provide strong evidences that the soluble active and the particulate latent are different forms of PPO in loquat fruit flesh. The results suggest that the major PPO form for the oxidation of CA, leading to enzymatic browning under physiological conditions, is the latent one.  相似文献   

15.
A kinetic framework is developed to describe enzyme activity and stability in two-phase liquid-liquid systems. In particular, the model is applied to the enzymatic production of benzaldehyde from mandelonitrile by Prunus amygdalus hydroxynitrile lyase (pa-Hnl) adsorbed at the diisopropyl ether (DIPE)/aqueous buffer interface (pH = 5.5). We quantitatively describe our previously obtained experimental kinetic results (Hickel et al., 1999; 2001), and we successfully account for the aqueous-phase enzyme concentration dependence of product formation rates and the observed reaction rates at early times. Multilayer growth explains the early time reversibility of enzyme adsorption at the DIPE/buffer interface observed by both enzyme-activity and dynamic-interfacial-tension washout experiments that replace the aqueous enzyme solution with a buffer solution. The postulated explanation for the unusual stability of pa-Hnl adsorbed at the DIPE/buffer interface is attributed to a two-layer adsorption mechanism. In the first layer, slow conformational change from the native state leads to irreversible attachment and partial loss of catalytic activity. In the second layer, pa-Hnl is reversibly adsorbed without loss in catalytic activity. The measured catalytic activity is the combined effect of the deactivation kinetics of the first layer and of the adsorption kinetics of each layer. For the specific case of pa-Hnl adsorbed at the DIPE/buffer interface, this combined effect is nearly constant for several hours resulting in no apparent loss of catalytic activity. Our proposed kinetic model can be extended to other interfacially active enzymes and other organic solvents. Finally, we indicate how interfacial-tension lag times provide a powerful tool for rational solvent selection and enzyme engineering.  相似文献   

16.
The hysteretic behavior of phosphoenolpyruvate (PEP) carboxylase from Crassula argentea has been investigated. Incubation of the purified enzyme with the inhibitor malate prior to starting the reaction by the addition of PEP resulted in a kinetic lag of several minutes duration. The length of the lag was inversely proportional to the enzyme concentration, suggesting subunit association-dissociation as the hysteretic mechanism, rather than a mechanism based on a slow conformational change in the enzyme. Dynamic laser light scattering measurements also support this conclusion, showing that the diffusion coefficient of malate-incubated enzyme slowly decreased after the reaction was started by the addition of PEP. Lags were observed only at pH values of 7.5 or lower. Maximum lags were observed after 10 min of preincubation with malate. Fumarate and succinate, which like malate caused mixed inhibition, also caused lags. In contrast, no lag was induced by malate in the presence of PEP or by the competitive inhibitor phosphoglycolate. The activators glucose 6-phosphate and malonate decreased the malate-induced lag.  相似文献   

17.
Polyphenol oxidase (PPO) was partially purified from dill by (NH4)(2)SO4 precipitation followed by dialysis and gel filtration chromatography. Polyphenol oxidase activity was measured spectrophotometrically at 420 nm using catechol, dopamine and chlorogenic acid as substrates. Optimum pH, temperature, and ionic strength were determined with three substrates. The best substrate of dill PPO was found to be chlorogenic acid. Some kinetic properties of the enzyme such as V(max,) K(M) and V(max)/K(M) were determined for all three substrates. The effects of various inhibitors on the reaction catalysed by the enzyme were tested and I(50) values calculated. The most effective inhibitor was L-cysteine. Activation energies, E(a), were determined from the Arrhenius equation. In addition, activation enthalpy, DeltaH(a), and Q(10) values of the enzyme were also calculated.  相似文献   

18.
During denaturation by sodium dodecyl sulfate (SDS), aminoacylase shows a rapid decrease in activity with increasing concentration of the detergent to reach complete inactivation at 1.0 mM SDS. The denatured minus native-enzyme difference spectrum showed two negative peaks at 287 and 295 nm. With the increase of concentration of SDS, both negative peaks increased in magnitude to reach maximal values at 5.0 mM SDS. The fluorescence emission intensity of the enzyme decreased, whereas there was no red shift of emission maximum in SDS solutions of increasing concentration. In the SDS concentration regions employed in the present study, no marked changes of secondary structure of the enzyme have been observed by following the changes in far-ultraviolet CD spectra. The inactivation of this enzyme has been followed and compared with the unfolding observed during denaturation in SDS solutions. A marked inactivation is already evident at low SDS concentration before significant conformational changes can be detected by ultraviolet absorbance and fluorescence changes. The inactivation rate constants of free enzyme and substrate-enzyme complex were determined by the kinetics method of the substrate reaction in the presence of inactivator previously described by Tsou [Tsou (1988),Adv. Enzymol. Related Areas Mol. Biol. 61, 381–436]. It was found that substrate protects against inactivation and at the same SDS concentrations, the inactivation rate of the free enzyme is much higher than the unfolding rate. The above results show that the active sites of metal enzyme containing Zn2+ are also situated in a limited and flexible region of the enzyme molecule that is more fragile to denaturants than the protein as a whole.  相似文献   

19.
The lactate dehydrogenase from Streptococcus faecalis is activated either by fructose 1,6-bisphosphate or by divalent cations such as Mn2+ or Co2+. With both types of activator, a lag is observed before attainment of the steady state rate of pyruvate reduction if the activator is added to the enzyme at the same time as the substrates. This lag can be largely abolished by preincubation of enzyme with activator before mixing with substrates. For fructose 1,6-bisphosphate (Fru(1,6)P2) as the activator, the rate constant for the lag phase showed a linear dependence on activator concentration but was independent of enzyme concentration. This suggests that binding of fructose 1,6-bisphosphate induces a conformational change in the enzyme which leads to increased activity, without association of enzyme subunits or dimers. With Co2+ as activator, the rate constant for the lag phase showed a hyperbolic dependence on Co2+ concentration and was also dependent on enzyme concentration. This suggests that activation by Co2+, in contrast to that by Fru(1,6)P2, involves association of enzyme dimers, followed by ligand binding.  相似文献   

20.
A highly active soluble peroxidase (donor: H2O2 oxidoreductase EC 1.11.1.7) has been purified from the preputial gland of the rat by hydroxylapatite chromatography, ammonium sulfate fractionation, Sephadex gel filtration and affinity chromatography on con A-Sepharose. The enzyme shows apparent homogeneity when analysed by acid and alkaline-PAGE. Its molecular, spectral, kinetic and catalytic properties were compared with those of bovine lactoperoxidase (LPO). Preputial gland peroxidase (PPO) is a glycoprotein of molecular weight of 70-73 kDa slightly lower (78 kDa) than that of LPO. Using isoelectric focussing, PPO was resolved into eight different closely spaced protein species spanning a pI range of 5.4 to 6.4, while LPO focuses into several closely spaced protein bands in the pI range 8.5-9.3. PPO is similar to LPO in its spectral (Soret) and some kinetic properties, but it differs significantly from LPO in substrate (H2O2) tolerance and substrate inactivation. PPO also differs from LPO in showing differential inactivation by SDS. Both enzymes are glycoproteins and although concanavalin A (con A) showed a variable interaction with both enzymes, wheat germ agglutinin interacted specifically with LPO only. We suggest that PPO, the secretory peroxidase of the preputial gland, differs significantly from LPO in its molecular and catalytic properties.  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