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.
The maturation of mastoparan B, the major toxin peptide in the venom of Vespa basalis, requires enzymatic cleavage of its prosequence presumably via sequential liberation of dipeptides. The putative processing enzyme, dipeptidyl peptidase IV, was expressed as a glycosylated His-tag fusion protein (rDPP-IV) via the baculovirus expression system. rDPP-IV purified by one-step nickel-affinity chromatography was verified by Western blot and LC-MS/MS analysis. The kcat/Km of rDPP-IV was determined to be in the range of 10–500 mM?1·S?1 for five synthetic substrates. The optimal temperature and pH for rDPP-IV were determined to be 50 °C and pH 9. Enzymatic activity of rDPP-IV was significantly reduced by 80 and 60% in the presence of sitagliptin and phenylmethylsulfonyl fluoride respectively.  相似文献   

2.
The covalent immobilization of bovine liver catalase (CAT) on to florisil via glutaraldehyde was investigated. Optimum immobilization pH and temperature were determined as pH 6.0, 10 degrees C respectively, while the amount of initial CAT per g of carrier and immobilization time was determined as 5 mg g(-1) and 120 min, respectively. The Vmax values for free and immobilized CAT were found to be 1.7 x 10(5) and 2.0 x 10(4) micromol H2O2 min(-1) mg protein(-1), respectively, whereas KM values were 33.3 mM and 1722.0 mM respectively. Operational stability was determined by using a stirred batch-type column reactor. Immobilized CAT retained about 40% of its initial activity after 50 uses. It showed higher storage stability than free CAT at 4 degrees C and 25 degrees C. Its storage stability increased with increasing relative humidity (RH) from 0 to 20% of the medium. The highest storage stability was obtained in 20% RH, however, further increase in RH from 40 to 100% significantly decreased the storage stability.  相似文献   

3.
Hemoglobin released in the circulation from ruptured red blood cells can be oxidized by hydrogen peroxide or peroxynitrite to generate the highly oxidizing iron(IV)oxo species HbFe(IV)z=O. Nitrogen monoxide, produced in large amounts by activated inducible nitric oxide synthase, can have indirect cytotoxic effects, mainly through the generation of peroxynitrite from its very fast reaction with superoxide. In the present work we have determined the rate constant for the reaction of HbFe(IV)z=O with NO(*), 2.4 x 10(7) M(-1)s(-1) at pH 7.0 and 20 degrees C. The reaction proceeds via the intermediate HbFe(III)ONO, which then dissociates to metHb and nitrite. As these products are not oxidizing and because of its large rate, the reaction of HbFe(IV)z=O with NO(*) may be important to remove the high valent form of hemoglobin, which has been proposed to be at least in part responsible for oxidative lesions. In addition, we have determined that the rate constant for the reaction of HbFe(IV)z=O with nitrite is significantly lower (7.5 x 10(2) M(-1)s(-1) at pH 7.0 and 20 degrees C), but increases with decreasing pH (1.8 x 10(3) M(-1)s(-1) at pH 6.4 and 20 degrees C). Thus, under acidic conditions as found in ischemic tissues, this reaction may also have a physiological relevance.  相似文献   

4.
The oxidation of serotonin (5-hydroxytryptamine) by the myeloperoxidase intermediates compounds I and II was investigated by using transient-state spectral and kinetic measurements at 25.0 +/- 0.1 degrees C. Rapid scan spectra demonstrated that both compound I and compound II oxidize serotonin via one-electron processes. Rate constants for these reactions were determined using both sequential-mixing and single-mixing stopped-flow techniques. The second order rate constant obtained for the one-electron reduction of compound I to compound II by serotonin is (1.7 +/- 0.1) x 10(7) M(-1) x s(-1), and that for compound II reduction to native enzyme is (1.4 +/- 0.1) x 10(6) M(-1) x s(-1) at pH 7.0. The maximum pH of the compound I reaction with serotonin occurs in the pH range 7.0-7.5. At neutral pH, the rate constant for myeloperoxidase compound I reacting with serotonin is an order of magnitude larger than for its reaction with chloride, (2.2 +/- 0.2) x 10(6) M(-1) x s(-1). A direct competition of serotonin with chloride for myeloperoxidase compound I oxidation was observed. Our results suggest that serotonin may have a role to protect lipoproteins from oxidation and to prevent enzymes from inactivation caused by the potent oxidants HOCl and active oxygen species.  相似文献   

5.
The regulation of intracellular pH during intense muscle contractions occurs via a number of different transport systems [e.g., monocarboxylate transporters (MCTs)] and via intracellular buffering (beta m(in vitro)). The aim of this study was to investigate the effects of an acute bout of high-intensity exercise on both MCT relative abundance and beta m(in vitro) in humans. Six active women volunteered for this study. Biopsies of the vastus lateralis were obtained at rest and immediately after 45 s of exercise at 200% of maximum O2 uptake. Beta m(in vitro) was determined by titration, and MCT relative abundance was determined in membrane preparations by Western blots. High-intensity exercise was associated with a significant decrease in both MCT1 (-24%) and MCT4 (-26%) and a decrease in beta m(in vitro) (-11%; 135 +/- 3 to 120 +/- 2 micromol H+ x g dry muscle(-1) x pH(-1); P < 0.05). These changes were consistently observed in all subjects, and there was a significant correlation between changes in MCT1 and MCT4 relative abundance (R2 = 0.92; P < 0.05). In conclusion, a single bout of high-intensity exercise decreased both MCT relative abundance in membrane preparations and beta m(in vitro). Until the time course of these changes has been established, researchers should consider the possibility that observed training-induced changes in MCT and beta m(in vitro) may be influenced by the acute effects of the last exercise bout, if the biopsy is taken soon after the completion of the training program. The implications that these findings have for lactate (and H+) transport following acute, exhaustive exercise warrant further investigation.  相似文献   

6.
N-carboxymethanofuran (carbamate) formation from unprotonated methanofuran (MFR) and CO2 is the first reaction in the reduction of CO2 to methane in methanogenic archaea. The reaction proceeds spontaneously. We address here the question whether the rate of spontaneous carbamate formation is high enough to account for the observed rate of methanogenesis from CO2. The rates of carbamate formation (v1) and cleavage (v2) were determined under equilibrium conditions via 2D proton exchange NMR spectroscopy (EXSY). At pH 7.0 and 300 K the second order rate constant k1* of carbamate formation from 'MFR'(MFR + MFRH+) and 'CO2' (CO2 + H2CO3 + HCO3-+ CO32-) was found to be 7 M-1.s-1 (v1 = k1* ['MFR'] ['CO2']) while the pseudo first order rate constant k2* of carbamate cleavage was 12 s-1 (v2 = k2* [carbamate]). The equilibrium constant K* = k1*/k2* = [carbamate]/['MFR']['CO2'] was 0.6 M-1 at pH 7.0 corresponding to a free energy change DeltaG degrees ' of + 1.3 kJ.mol-1. The pH and temperature dependence of k1*, of k2* and of K* were determined. From the second order rate constant k1* it was calculated that under physiological conditions the rate of spontaneous carbamate formation is of the same order as the maximal rate of methane formation and as the rate of spontaneous CO2 formation from HCO3- in methanogenic archaea, the latter being important as CO2 is mainly present as HCO3- which has to be converted to CO2 before it can react with MFR. An enzyme catalyzed carbamate formation thus appears not to be required for methanogenesis from CO2. Consistent with this conclusion is our finding that the rate of carbamate formation was not enhanced by cell extracts of Methanosarcina barkeri and Methanobacterium thermoautotrophicum or by purified formylmethanofuran dehydrogenase which catalyzes the reduction of N-carboxymethanofuran to N-formylmethanofuran. From the concentrations of 'CO2' and of 'MFR' determined by 1D-NMR spectroscopy and the pKa of H2CO3 and of MFRH+ the concentrations of CO2 and of MFR were obtained, allowing to calculate k1 (v1 = k1 [MFR] [CO2]). The second order rate constant k1 was found to be approximately 1000 M-1 x s-1 at 300 K and pH values between 7.0 and 8. 0 which is in the order of k1 values determined for other carbamate forming reactions by stopped flow.  相似文献   

7.
The effects of external pH on the efflux of protons from illuminated spinach chloroplasts have been studied by monitoring the rates of proton-pumping electron transport under a variety of steady-state conditions. Phosphorylation-coupled proton efflux through the ATP synthase (CF0-CF1), determined from the rates of ATP formation and that portion of the total electron transport attributable to phosphorylation, is strongly dependent upon pH over the range 6–9, with little activity below pH 7 and half-maximal activity at pH ≈ 7.6. Noncoupled proton efflux through the ATP synthase, determined in the absence of ADP and phosphate, was also strongly pH sensitive, with little activity below pH 7.5 and half-maximal activity at pH ~- 7.9. When proton efflux via CF0 was prevented by triphenyltin, the rate of passive proton leakage across the membrane was very low and practically insensitive to external pH indicating that the major pH-sensitive pathway(s) for proton efflux in the light involves CF0 · CF1. Modification of CF1 sulfhydryls by Ag+ resulted in an apparent increase in proton efflux via the normally coupled CF0 · CF1 pathway (half-maximal activity = pH 7.6), whereas modification by Hg2+ resulted in an apparent increase in proton efflux via the noncoupled CF0 · CF1 pathway (half-maximal activity = pH 7.9).  相似文献   

8.
The cytoplasmic domain of the anion exchange protein (cdb3) serves as a critical organizing center for protein-protein interactions that stabilize the erythrocyte membrane. The structure of the central core of cdb3, determined by X-ray crystallography from crystals grown at pH 4.8, revealed a compact dimer for residues 55-356 and unresolved N- and C-termini on each monomer [Zhang et al. (2000) Blood 96, 2925-2933]. Given that previous studies had suggested a highly asymmetric structure for cdb3 and that pH dependent structural transitions of cdb3 have been reported, the structure of cdb3 in solution at neutral pH was investigated via site-directed spin labeling in combination with conventional electron paramagnetic resonance (EPR) and double electron electron resonance (DEER) spectroscopies. These studies show that the structure of the central compact dimer (residues 55-356) is indistinguishable from the crystal structure determined at pH 4.8. N-Terminal residues 1-54 and C-terminal residues 357-379 are dynamically disordered and show no indications of stable secondary structure. These results establish a structural model for cdb3 in solution at neutral pH which represents an important next step in characterizing structural details of the protein-protein interactions that stabilize the erythrocyte membrane.  相似文献   

9.
A novel pH-sensitive amphiphilic copolymer brush poly(methyl methacrylate-co-methacrylic acid)-b-poly(poly(ethylene glycol) methyl ether monomethacrylate) [P(MMA-co-MAA)-b-PPEGMA] was defined and synthesized by atom transfer radical polymerization (ATRP) technique. The molecular structures and characteristics of this copolymer and its precursors were confirmed by (1)H NMR, FT-IR, and GPC. The CMC of P(MMA-co-MAA)-b-PPEGMA in aqueous medium was determined to be 1-4 mg/L. This copolymer could self-assemble into micelles in aqueous solution with an average size of 120-250 nm determined by DLS. The morphologies of the micelles were found to be spherical by SEM and TEM. Ibuprofen (IBU), a poorly water-soluble drug, was selected as the model drug and wrapped into the core of micelles via dialysis method. Drug entrapment efficiency reached to 90%. The in vitro release behavior of IBU from these micelles was pH-dependent. The cumulative release percent of IBU was less than 20% of the initial drug content in simulated gastric fluid (SGF, pH 1.2) over 12 h, but 90% was released in simulated intestinal fluid (SIF, pH 7.4) within 6 h. The release profiles showed that the P(MMA-co-MAA)-b-PPEGMA micelles could inhibit the premature burst drug release under the intestinal conditions. All the results indicate that the P(MMA-co-MAA)-b-PPEGMA micelle may be a potential oral drug delivery carrier for poorly water-soluble drugs.  相似文献   

10.
The ability to regulate intracellular pH (pH(i)) is essential for normal cell development and differentiation. This study was an investigation of the regulatory system used by the hamster oocyte and preimplantation embryo to regulate pH(i) in the alkaline range. Recovery from alkalosis by late 1-cell and 2-cell embryos was rapid, and physiological pH(i) levels could be restored within 10 min. Recovery from an induced alkaline load was dependent on the chloride concentration in the external medium and sensitive to a stilbene derivative 4,4'-diisothiocyanatostilbene-2,2'-di-sulfonic acid that inhibits bicarbonate and chloride exchange. Therefore the recovery from alkalosis by hamster embryos appears to be via activity of the HCO(3)(-)/Cl(-) exchanger that was activated above a pH(i) set point of 7.24. In contrast, hamster oocytes and early 1-cell embryos (collected 3-4 h post-egg activation) could not recover from an intracellular alkalosis, and pH(i) remained elevated. Therefore, the hamster oocyte and the early 1-cell embryo still undergoing pronuclear formation lack an active HCO(3)(-)/Cl(-) exchanger for the restoration of pH(i). Inability to restore pH(i) from an alkali challenge resulted in a reduced ability of embryos to develop to the morula/blastocyst stages in culture, indicating that HCO(3)(-)/Cl(-) exchange is involved in physiological regulation of pH(i).  相似文献   

11.
Herold S  Exner M  Nauser T 《Biochemistry》2001,40(11):3385-3395
The second-order rate constants for the reactions between nitrogen monoxide and oxymyoglobin or oxyhemoglobin, determined by stopped-flow spectroscopy, increase with increasing pH. At pH 7.0 the rates are (43.6 +/- 0.5) x 10(6) M(-1) x s(-1) for oxymyoglobin and (89 +/- 3) x 10(6) M(-1) x s(-1) for oxyhemoglobin (per heme), whereas at pH 9.5 they are (97 +/- 3) x 10(6) M(-1) x s(-1) and (144 +/- 3) x 10(6) M(-1) x s(-1), respectively. The rate constants for the reaction between oxyhemoglobin and NO* depend neither on the association grade of the protein (dimer/tetramer) nor on the concentration of the phosphate buffer (100-1 mM). The nitrogen monoxide-mediated oxidations of oxymyoglobin and oxyhemoglobin proceed via intermediate peroxynitrito complexes which were characterized by rapid scan UV/vis spectroscopy. The two complexes MbFe(III)OONO and HbFe(III)OONO display very similar spectra with absorption maxima around 500 and 635 nm. These species can be observed at alkaline pH but rapidly decay to the met-form of the proteins under neutral or acidic conditions. The rate of decay of MbFe(III)OONO increases with decreasing pH and is significantly larger than those of the analogous complexes of the two subunits of hemoglobin. No free peroxynitrite is formed during these reactions, and nitrate is formed quantitatively, at both pH 7.0 and 9.0. This result indicates that, as confirmed from protein analysis after reacting the proteins with NO* for 10 times, when peroxynitrite is coordinated to the heme of myoglobin or hemoglobin it rapidly isomerizes to nitrate without nitrating the globins in physiologically significant amounts.  相似文献   

12.
Lin SY  Hsieh TF  Wei YS 《Peptides》2005,26(4):543-549
The secondary structure of PGAIPG (Pro-Gly-Ala-IIe-Pro-Gly), a repeated hexapeptide of tropoelastin, in buffer solution of different pH was determined by using attenuated total reflection-Fourier transform infrared (ATR-FTIR) spectroscopy. The thermal-dependent structural change of PGAIPG in aqueous solution or in solid state was also examined by thermal FTIR microspectroscopy. The conformation of PGAIPG in aqueous solution exhibited a pH-dependent structural characterization. A predominant peak at 1614 cm(-1) (aggregated beta-sheet) with a shoulder near 1560 cm(-1) (beta-sheet) appeared in pH 5.5-8.5 buffer solutions. A new broad shoulder at 1651 cm(-1) (random coil and/or alpha-helix) with 1614 cm(-1) was observed in the pH 4.5 buffer solution. However, the broad shoulder at 1651 cm(-1) was converted to a maximum peak at 1679 cm(-1) (beta-turn/antiparallel beta-sheet) when the pH shifted from 4.5 to 3.5, but the original pronounced peak at 1614 cm(-1) became a shoulder. Once the pH was lowered to 2.5, the IR spectrum of PGAIPG was dominated by major absorption at 1679 cm(-1) with a minor peak at 1552 cm(-1) (alpha-helix/random coil). The result indicates that the pH was a predominant factor to transform PGAIPG structure from aggregated beta-sheet (pH 8.5) to beta-turn/intermolecular antiparallel beta-sheet (pH 2.5). Moreover, a partial conformation of PGAIPG with minor alpha-helix/random coil structures was also explored in the lower pH buffer solution. There was no thermal-dependent structural change for solid-state PGAIPG. The thermal-induced formation of aggregated beta-sheet for PGAIPG in aqueous solution was found from 28 to 30 degrees C, however, which might be correlated with the formation of an opaque gel that turned from clear solution. The formation of aggregated beta-sheet structure for PGAIPG beyond 30 degrees C might be due to the intermolecular hydrogen bonded interaction between the hydrophobic PGAIPG fragments induced by coacervation.  相似文献   

13.
The manganese meso-dimethylimidazolium porphyrin complex Mn(III)[TDMImP] reacted with HOBr/OBr(-) to generate the corresponding oxo-Mn(V)[TDMImP] species. The rate of this process accelerated with increasing pH. A forward rate constant, k(for), of 1.65x10(6)M(-1)s(-1) was determined at pH 8. Under these conditions, the oxo-Mn(V) species is short-lived and is transformed into the corresponding oxo-Mn(IV) complex. A first-order rate constant, k(obs), of 0.66 s(-1) was found for this reduction process at pH 8. The mechanism of this reduction process, which was dependent on bromide ion, appeared to proceed via an intermediate Mn(III)-O-Br complex. Thus, both a fast, reversible Mn(III)-O-Br bond heterolysis and a slower homolytic pathway occur in parallel in this system. The reverse oxidation reaction between oxo-Mn(V)[TDMImP] and bromide was investigated as a function of pH. The rate of this oxo-transfer reaction (k(rev)=1.4x10(3)M(-1)s(-1) at pH 8) markedly accelerated as the pH was lowered. The observed first-order dependence of the rate on [H(+)] indicates that the reactive species responsible for bromide oxidation is a protonated oxo-hydroxo complex and the stable species present in solution at high pH is dioxo-Mn(V)[TDMImP], [O=Mn(V)=O](-). The oxo-Mn(V) species retains nearly all of the oxidative driving force of the hypohalite. The equilibrium constant K(equi)=k(for)/k(rev) for the reversible process was determined at three different pH values (K(equi)=1.15x10(3) at pH 8) allowing the measurement of the redox potentials E of oxo-Mn(V)/Mn(III) (E=1.01 V at pH 8). The redox potential for this couple was extrapolated over the entire pH scale using the Nernst relationship and compared to those of the manganese 2- and 4-meso-N-methylpyridinium porphyrin couples oxo-Mn(V)[2-TMPyP]/Mn(III)[2-TMPyP], oxo-Mn(V)[4-TMPyP]/Mn(III)[4-TMPyP], OBr(-)/Br(-) and H(2)O(2)/H(2)O. Notably, the redox potential of oxo-Mn(V)/Mn(III) for the imidazolium porphyrin approaches that of H(2)O(2)/H(2)O at low pH.  相似文献   

14.
Cyclic maltosyl-maltose [CMM, cyclo-[-->6)-alpha-D-Glcp-(1-->4)-alpha-D-Glcp-(1-->6)-alpha-D-Glcp-(1-->4)-alpha-D-Glcp-(1-->]], a novel cyclic tetrasaccharide, has a unique structure. Its four glucose residues are joined by alternate alpha-1,4 and alpha-1,6 linkages. CMM is synthesized from starch by the action of 6-alpha-maltosyltransferase from Arthrobacter globiformis M6. Recently, we determined the mechanism of extracellular synthesis of CMM, but the degrading pathway of the saccharide remains unknown. Hence we tried to identify the enzymes involved in the degradation of CMM to glucose from the cell-free extract of the strain, and identified CMM hydrolase (CMMase) and alpha-glucosidase as the responsible enzymes. The molecular mass of CMMase was determined to be 48.6 kDa by sodium dodecyl sulfate polyacrylamide gel electrophoresis (SDS-PAGE), and 136 kDa by gel filtration column chromatography. The optimal pH and temperature for CMMase activity were 6.5 and 30 degrees C. The enzyme remained stable from pH 5.5 to 8.0 and up to 25 degrees C. CMMase hydrolyzed CMM to maltose via maltosyl-maltose as intermediates, but it did not hydrolyze CMM to glucose, suggesting that it is a novel hydrolase that hydrolyzes the alpha-1,6-linkage of CMM. The molecular mass of alpha-glucosidase was determined to be 60.1 kDa by SDS-PAGE and 69.5 kDa by gel filtration column chromatography. The optimal pH and temperature for alpha-glucosidase activity were 7.0 and 35 degrees C. The enzyme remained stable from pH 7.0 to 9.5 and up to 35 degrees C. alpha-Glucosidase degraded maltosyl-maltose to glucose via panose and maltose as intermediates, but it did not degrade CMM. Furthermore, when CMMase and alpha-glucosidase existed simultaneously in a reaction mixture containing CMM, glucose was detected as the final product. It was found that CMM was degraded to glucose by the synergistic action of CMMase and alpha-glucosidase.  相似文献   

15.
The kinetics of the recombination of the metal-depleted active site of horse liver alcohol dehydrogenase (LADH) with metal ions have been studied over a range of pH and temperature. The formation rates were determined optically, by activity measurements, or by using the pH change during metal incorporation with a pH-indicator as monitor. The binding of Zn2+, Co2+, and Ni2+ ions occurs in a two-step process. The first step is a fast equilibrium reaction, characterized by an equilibrium constant K1. The spectroscopic and catalytic properties of the native or metal-substituted protein are recovered in a slow, monomolecular process with the rate constant k2. The rate constants k2 5.2 X 10(-2) sec-1 (Zn2+), 1.1 X 10(-3) sec-1 (Co2+), and 2 X 10(-4) sec-1 (Ni2+). The rate constants increase with increasing pH. Using temperature dependence, the activation parameters for the reaction with Co2+ and Ni2+ were determined. Activation energies of 51 +/- 2.5 kJ/mol (0.033 M N-Tris-(hydroxymethyl)methyl-2-aminomethane sulfonic acid (TES), pH 6, 9) for Co2+ and 48.5 +/- 4 kJ/mol (0.033 M TES, pH 7, 2) for Ni2+ at 23 degrees C were found. The correspondent activation entropies are - 146 +/- 10 kJ/mol K for Co2+ and - 163 +/- 9 kJ/mol K for Ni2+. Two protons are released during the binding of Zn2+ to H4Zn(n)2 LADH in the pH range 6.8-8.1. The binding of coenzyme, either reduced or oxidized, prevents completely the incorporation of metal ions, suggesting that the metal ions enter the catalytic site via the coenzyme binding domain and not through the hydrophobic substrate channel.  相似文献   

16.
Wijma HJ  Canters GW  de Vries S  Verbeet MP 《Biochemistry》2004,43(32):10467-10474
The copper-containing nitrite reductase from Alcaligenes faecalis S-6 was found to catalyze the oxidation of nitric oxide to nitrite, the reverse of its physiological reaction. Thermodynamic and kinetic constants with the physiological electron donor pseudoazurin were determined for both directions of the catalyzed reaction in the pH range of 6-8. For this, nitric oxide was monitored by a Clark-type electrode, and the redox state of pseudoazurin was measured by optical spectroscopy. The equilibrium constant (K(eq)) depends on the reduction potentials of pseudoazurin and nitrite/nitric oxide, both of which vary with pH. Above pH 6.2 the formation of NiR substrates (nitrite and reduced pseudoazurin) is favored over the products (NO and oxidized pseudoazurin). At pH 8 the K(eq) amounts to 10(3). The results show that dissimilatory nitrite reductases catalyze an unfavorable reaction at physiological pH (pH = 7-8). Consequently, nitrous oxide production by copper-containing nitrite reductases is unlikely to occur in vivo with a native electron donor. With increasing pH, the rate and specificity constant of the forward reaction decrease and become lower than the rate of the reverse reaction. The opposite occurs for the rate of the reverse reaction; thus the catalytic bias for nitrite reduction decreases. At pH 6.0 the k(cat) for nitrite reduction was determined to be 1.5 x 10(3) s(-1), and at pH 8 the rate of the reverse reaction is 125 s(-1).  相似文献   

17.
Phospholipase C-gamma1 displayed sigmoidal kinetics with a S(0.5) value of 0.17 mole fraction PIP(2) when assayed at pH 6.8 using detergent:lipid mixed micelles. The pH optimum for hydrolysis of phosphatidylinositol 4,5-bisphosphate by phospholipase C-gamma1 was dependent on the mole fraction of substrate in the micelle. The pH optimum was 5.5 when the enzyme was assayed below the S(0.5). The pH optima shifted to a pH range of 6.0-6.3 when the enzyme was assayed above the S(0.5). The kinetic parameters for phospholipase C-gamma1 assayed at various pH values from pH 7.0 to 5.0 yielded similar n values (n=4), but the constant, K', decreased from 1x10(-2) (mole fraction)(2) at pH 7.0 to 1x10(-5) (mole fraction)(2) at pH 5.0. Maximum enzyme specificity occurred at pH values below pH 6.0 as determined by the plot of logk(cat)/S(0.5) versus pH. Intrinsic fluorescence spectroscopy revealed that at a pH value above 7.0 or below 6.3, tryptophan quenching occurred. Fluorescence quenching experiments performed with acrylamide determined phospholipase C-gamma1 incubated at pH 5.0 had a larger collisional quenching constant than enzyme incubated at pH 7.0. Lowering the pH to 5.0 apparently resulted in interior tryptophans becoming more solvent accessible. These data suggest that pH may activate phospholipase C-gamma1 by disrupting ionizable groups leading to a conformational change.  相似文献   

18.
A thermostable monoacylglycerol lipase [MGLP, EC 3.1.1.23] was purified for the first time from a cell-free extract of the moderately thermophilic Bacillus sp. H-257. The enzyme was purified 3,028-fold to homogeneity by chromatography using Octyl-Sepharose CL-4B, Q-Sepharose FF, and Superose 12 columns. The molecular mass of the MGLP was estimated to be 25 kDa by gel filtration and 24 kDa by SDS-PAGE, suggesting a monomeric protein. The isoelectric point was determined to be 4.66 by isoelectric focusing. The MGLP retained its full activity upon incubation at 60 degrees C for 10 min (pH 7. 3), and was stable at pH 7-10. The optimal temperature for activity at pH 7.5 was 75 degrees C, and the maximum activity was observed from pH 6-8. This enzyme hydrolyzes monoacylglycerols, with the highest activity occurring with 1-monolauroylglycerol. Di- and triacylglycerols, on the other hand, are essentially inert as substrates for the enzyme. The K(m) values for the hydrolysis of 1-monolauroylglycerol, 1-monooleoylglycerol, and 2-monooleoylglycerol were determined to be 140, 83 and 59 mM, respectively. The enzyme was not inhibited by cholate, but was slightly inhibited by Triton X-100 and deoxycholate. The amino acid sequence of the N-terminal region of the enzyme (16 residues) was also determined.  相似文献   

19.
H Yoshida  H Hanazawa 《Biochimie》1989,71(5):687-692
Ribonuclease (RNase) F1 was inactivated by incubation with an excess amount of iodoacetate at pH 5.5, 37 degrees C according to pseudo first-order kinetics. It was protected to various degrees, from inactivation by nucleotides, among which guanosine 2'-phosphate was most effective. The pseudo first-order rate constant was proportional to the reagent concentration, indicating that the reaction in reality follows second-order kinetics. The second-order rate constant was determined to be 25 x 10(-4) M-1 s-1. The inactivation rate was maximal at pH 5.5-6.0. When iodo[2-14C]acetate was used as the reagent, the stoichiometry of incorporation was determined to be 1.1 mol carboxymethyl group per mol of RNase F1 and glutamic acid residue 58 was assigned as the site of modification.  相似文献   

20.
The viability of Clavibacter michiganensis subsp. michiganensis (Cmm) was determined by measuring the intracellular pH (pHin) as a viability parameter. This was based on the observation that growth of Cmm was inhibited at pH 5.5 and below. Therefore, viable cells should maintain their pHin above this pH value. The pHin of Cmm was determined using the fluorescent probe 5(and 6-)-carboxyfluorescein succinimidyl ester (cFSE). The pHin of Cmm cells exposed to acid treatments was determined using fluorescence spectrofluorometry, and for cells exposed to elevated temperatures, the pHin was determined using fluorescence spectrofluorometry and flow cytometry (FCM). A good correlation was found between the presence of a pH gradient and the number of colony-forming units (cfu) observed in plate counts. However, with the spectrofluorometry technique, the analysis is based on the whole cell population and the detection sensitivity of this technique is rather low, i.e., cell numbers of at least 107 cfu ml-1 are needed for the analysis. Using FCM, heat-treated and non-treated Cmm cells could be distinguished based on the absence and presence of a pH gradient, respectively. The major advantage of FCM is its high sensitivity, allowing analysis of microbial populations even at low numbers, i.e., 102-103 cfu ml-1.  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