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.
Fructansucrases, members of glycoside hydrolase family 68, catalyze both sucrose hydrolysis and the polymerization of fructose to beta-d-fructofuranose polymers. The resulting fructan polymers are distinguished by the nature of the glycosidic bond: inulin (beta-(2-1)-fructofuranose) and levan (beta-(2-6)-fructofuranose). In this study we demonstrate that Zymomonas mobilis levansucrase exists in two active forms, depending on the pH and ionic strength. At pH values above 7.0, the enzyme is mainly a dimer, whereas at pH values below 6.0, the protein forms well ordered microfibrils that precipitate out of the solution. These two forms are readily interchangeable simply by changing the pH. Surprisingly the manner in which the enzyme is arranged strongly affects its product specificity and kinetic properties. At pH values above 7.0, the activity of the enzyme as a dimer is mainly sucrose hydrolysis and the synthesis of short fructosaccharides (degree of polymerization, 3). At pH values below 6.0, in its microfibril form, the enzyme catalyzes almost exclusively the synthesis of levan (a degree of polymerization greater than 20,000). This difference in product specificity appears to depend on the form of the enzyme, dimer versus microfibril, and not directly on the pH. Images made by negative stain transmission electron microscopy reveal that the enzyme forms a very ordered structure of long fibrils that appear to be composed of repeating rings of six to eight protein units. A single amino acid replacement of H296R abolished the ability of the enzyme to form microfibrils with organized fibril networks and to synthesize levan at pH 6.0.  相似文献   

2.
The decomposition kinetics of bis-POC PMEA and bis-POC PMPA followed pseudo-first order kinetics with the corresponding mono-POC ester detected as the only observable degradation product for all the pH values studied. The rates of hydrolysis of bis-POC PMEA over the pH range studied was described by [formula: see text] The 18O incorporation studies revealed that hydrolysis of bis-POC PMEA at pH 7.0 primarily proceeds via P-O cleavage with an additional minor pathway involving C-O bond cleavage. Hydrolysis of bis-POC PMPA was found to be about 2 fold slower than bis-POC PMEA at pH values above 6.0.  相似文献   

3.
Fixation with Bouin's fluid preserves cytoplasmic and nucleolar ribonucleic acid (UNA) particularly well. RNA may be demonstrated preferentially in Bouin fixed tissue by staining with 0.02% thiazine dye in aqueous McIIvaine phosphate-citrate buffer between pH 3 and 4. Methylation blockage of basophilia other than that of nucleic acids permits staining of RNA with thiazine dyes near neutrality. The deoxyribonucleic acid (DNA) of chromatin undergoes a Feulgen type hydrolysis in the tissue block during 24 hr fixation with Bouin's fluid. This hydrolysis by picric acid permits Schiff staining of the DNA wthout further acid hydrolysis. Consequently after Bouin fixation it is possible to demonstrate DNA and RNA specifically by a Schiff-methylene blue sequence. Thus a Schiff stain without further acid hydrolysis followed by 0.02% methylene blue in phosphate-citrate buffer at pH 3.0 to 3.5 colors DNA magenta in contrast to the blue of RNA.  相似文献   

4.
Over a pH range 1-4 and temperatures from 170 to 230 degrees C, the decomposition rates of xylose, galactose, mannose, glucose, 2-furfural, and 5-hydroxymethyl-2-furfural (5-HMF) were pseudo first order. The effect of temperature and pH on the pseudo first-order decomposition rate constants was modeled using the Arrhenius equation and acid-base catalysis, respectively. Decomposition rates of the monosaccharides were minimum at a pH 2-2.5. Above pH 2.5, the monosaccharide decomposition was base catalyzed, with acid catalysis occurring at a pH of less than 2 for glucose. The furfurals were subject to acid catalysis at below ca. pH 3.5. The hydrothermal conversion of glucose to its decomposition products during thermochemical Pretreatment can be modeled as a combination of series and parallel reactions. The formation rates of identified soluble products from glucose decomposition, 5-HMF and levulinic acid, were also functions of temperature and pH. The rate of 5-HMF formation relative to glucose decomposition decreased as the pH increased from 2.0 to 4.0, with levulinic acid formation only detected when the pH was 2.5 or less. For glucose decomposition, humic solids accounted for ca. 20% of the decomposition products.  相似文献   

5.
Generation of diastereomeric phosphonate ester adducts of chymotrypsin was evidenced for the first time by 31P NMR and spectrophotometric kinetic measurements. 31P NMR signals were recorded for 4-nitrophenyl 2-propyl methylphosphonate (IMN) at 32.2 ppm and for its hydrolysis product at 26.3 ppm downfield from phosphoric acid. The inhibition of α-chymotrypsin at pH > 8.0 by the faster reacting enantiomer of IMN or 2-propyl methylphosphonochloridate (IMCl), or other phosphonate ester analogs of these compounds, all caused a ~6.0 ppm downfield shift of the 31P signal to the 39–40 ppm region. IMN, when applied below the stoichiometric amount of chymotrypsin, under the same conditions, generated two signals, at 39.0 and at 37.4 ppm. Scans accumulated in hourly intervals showed the decomposition of both diastereomers, with approximate half-lives of 12 h at pH 8.0 and 22°C, into a species with a resonance at 35.5 ppm. The most likely reaction to account for the appearance of this new peak is the enzymic dealkylation of the isopropyl group from the covalently bound phosphonate ester. We base this conclusion mostly on the similarity of the upfield shift to the hydrolysis of phosphonate esters. Contrary to experience with phosphate ester adducts of serine proteases, no signal was detected higher than 25.0 ppm downfield from phosphoric acid for several phosphonate ester adducts of chymotrypsin and in no case did the resonance for the adduct shift further downfield in the course of the experiments. © 1993 Wiley-Liss, Inc.  相似文献   

6.
At pH 7.35, N-(2-oxopropyl)-N-nitrosourea (OPNU) reacted with calf thymus DNA to yield O6-methylguanine, 7-methylguanine and 3-methyladenine. Kinetic measurements of the base catalyzed decomposition of OPNU and the extent of methylation of DNA by OPNU suggested that methylnitrosourea is not formed as an intermediate product. Diazoacetone, acetic acid and methanol were identified as products of decomposition of OPNU at pH 7.35. Reaction of OPNU with N-methylmaleimide yielded the product resulting from 1,3-dipolar cycloaddition of diazomethane. Hydrolysis of N-nitroso-N-acetoxymethyl-N-2-oxopropylamine (NAMOPA) in the presence of hog liver esterase also produced diazoacetone, acetic acid and methanol. Enzymatic hydrolysis of NAMOPA in the presence of DNA produced O6-methylguanine, 7-methylguanine and 3-methyladenine. These results suggest that OPNU undergoes base-catalyzed decomposition and NAMOPA undergoes enzymatic hydrolysis to yield the same intermediate, 2-oxopropyldiazotate. This diazotate then reacts either by protonation followed by loss of water to form diazoacetone, or by internal nucleophilic attack by the diazotate oxygen on the carbonyl carbon to form an oxadiazoline intermediate which then collapses to form acetate and the methylating agent diazomethane. These reaction schemes are used to suggest the mechanism by which N-nitroso-2-oxopropylpropylamine methylates hepatic DNA in vivo.  相似文献   

7.
Catalytic properties (KM, Vmax) of aminopeptidase in pig kidney sections, in isolated membranes and in a solubilized purified form were investigated using amino acid 2-naphthylamides and 4-methoxy-2-naphthylamides. In the first case these properties were estimated on the basis of the stain intensity resulting from the coupling of product with Fast Blue B, in the latter two cases they were measured fluorometrically. The following observations were made: (1) In all three cases the substrate turnover was shown to be a direct function of time and enzyme concentration. (2) The values obtained for the solubilized and the membrane bound form were practically identical but differed from those found in tissue sections. (3) Each amino acid derivative had defined constants, but these were difficult to obtain in sections, especially if it was necessary, on account of poor solubilities, to use low substrate concentrations. (4) Hydrophilic amino acid derivatives were adsorbed to tissue membranes much less than hydrophobic ones. (5) Fast Blue B caused a non-competitive inhibition of enzymic activity. (6) Binding of antibody against pure aminopeptidase caused inhibition of the enzymic hydrolysis of all the naphthylamides. Thus, histochemical stain intensities per time and area derived from one substrate at a defined concentration are suitable for the determination of enzyme concentrations. However, no conclusions regarding the homogeneity of the enzyme in sections can be drawn by comparing the stain intensities obtained with different substrates in contrast to data from the inhibition of substrate hydrolysis by antibody.  相似文献   

8.
Summary Catalytic properties (KM, Vmax) of aminopeptidase in pig kidney sections, in isolated membranes and in a solubilized purified form were investigated using amino acid 2-naphthylamides and 4-methoxy-2-naphthylamides. In the first case these properties were estimated on the basis of the stain intensity resulting from the coupling of product with Fast Blue B, in the latter two cases they were measured fluorometrically. The following observations were made: (1) In all three cases the substrate turnover was shown to be a direct function of time and enzyme concentration. (2) The values obtained for the solubilized and the membrane bound form were practically identical but differed from those found in tissue sections. (3) Each amino acid derivative had defined constants, but these were difficult to obtain in sections, especially if it was necessary, on account of poor solubilities, to use low substrate concentrations. (4) Hydrophilic amino acid derivatives were adsorbed to tissue membranes much less than hydrophobic ones. (5) Fast Blue B caused a non-competitive inhibition of enzymic activity. (6) Binding of antibody against pure aminopeptidase caused inhibition of the enzymic hydrolysis of all the naphthylamides. Thus, histochemical stain intensities per time and area derived from one substrate at a defined concentration are suitable for the determination of enzyme concentrations. However, no conclusions regarding the homogeneity of the enzyme in sections can be drawn by comparing the stain intensities obtained with different substrates in contrast to data from the inhibition of substrate hydrolysis by antibody.  相似文献   

9.
The chemical nature of the phosphoryl enzyme linkage of the electrogenic proton-translocating ATPase (ATP phosphohydrolase, EC 3.6.1.3) in the plasma membrane of Neurospora has been identified as a mixed anhydride between phosphate and the beta-carboxyl group of an aspartic acid residue in the polypeptide chain. Incubation of isolated Neurospora plasma membrane vesicles containing 32P-labeled ATPase in buffers of increasing pH followed by analysis of the hydrolysis products yielded a pH versus hydrolysis profile characteristic of an acyl phosphate linkage. Reaction of labeled membranes with hydroxylamine at pH 5.3 also released [32P]i from the ATPase. Amino acid analyses of the Na[3H]BH4 reduction products obtained from membranes containing phosphorylated and dephosphorylated ATPase identified [3H]homoserine, the expected reduction product of beta-aspartyl phosphate, as the only additional tritiated reduction product in the samples from phosphorylated membranes. Tritium was not found in alpha-amino-delta-hydroxyvaleric acid, the reduction product of gamma-glutamyl phosphate, nor in proline, the degradation product of alpha-amino-delta-hydroxyvaleric acid. These results indicate that the phosphorylated intermediate of the Neurospora plasma membrane ATPase is a beta-aspartyl phosphate identical with that already known to exist in the Na+:K+- and Ca2+-translocating ATPases of animal cell origin. A common model for the mechanisms of all 3 ion-translocating ATPases is presented.  相似文献   

10.
A comparative study on the decomposition of Japanese red pine wood under subcritical water conditions in the presence and absence of phosphate buffer was investigated in a batch-type reaction vessel. Since cellulose makes up more than 40-45% of the components found in most wood species, a series of experiments were also carried out using pure cellulose as a model for woody biomass. Several parameters such as temperature and residence time, as well as pH effects, were investigated in detail. The best temperature for decomposition and hydrolysis of pure cellulose was found around 270 °C. The effects of the initial pH of the solution which ranged from 1.5 to 6.5 were studied. It was found that the pH has a considerable effect on the hydrolysis and decomposition of the cellulose. Several products in the aqueous phase were identified and quantified. The conditions obtained from the subcritical water treatment of pure cellulose were applied for the Japanese red pine wood chips. As a result, even in the absence of acid catalyst, a large amount of wood sample was hydrolyzed in water; however, by using phosphate buffer at pH 2, there was an increase in the hydrolysis and dissolution of the wood chips. In addition to the water-soluble phase, acetone-soluble and water-acetone-insoluble phases were also isolated after subcritical water treatment (which can be attributed mainly to the degraded lignin, tar, and unreacted wood chips, respectively). The initial wood:acid ratio in the case of reactions catalyzed by phosphate buffer was also investigated. The results showed that this weight ratio can be as high as 3:1 without changing the catalytic activity. The size of the wood chips as one of the most important experimental parameters was also investigated.  相似文献   

11.
Hyaluronic acid (HA) was hydrolyzed using varying temperatures (40, 60, and 80 degrees C) and acid concentrations (0.0010, 0.010, 0.10, 0.50, 1.0, and 2.0 M HCl). The degradation process was monitored by determination of weight average molecular weight ( M w) by size-exclusion chromatography with online multiangle laser light scattering, refractive index, and intrinsic viscosity detectors (SEC-MALLS-RI-visc) on samples taken out continuously during the hydrolysis. SEC-MALLS-RI-visc showed that the degradation gave narrow molecular weight distributions with polydispersity indexes ( M w/ M n) of 1.3-1.7. Kinetic plots of 1/ M w versus time gave linear plots showing that acid hydrolysis of HA is a random process and that it follows a first order kinetics. For hydrolysis in HCl at 60 and 80 degrees C, it was shown that the kinetic rate constant ( k h) for the degradation depended linearly on the acid concentration. Further, the dependence of temperature on the hydrolysis in 0.1 M HCl was found to give a linear Arrhenius plot (ln k h vs 1/ T), with an activation energy ( E a) of 137 kJ/mol and Arrhenius constant ( A) of 7.86 x 10 (15) h (-1). (1)H NMR spectroscopy was used to characterize the product of extensive hydrolysis (48 h at 60 degrees C in 0.1 M HCl). No indication of de- N-acetylation of the N-acetyl glucosamine (GlcNAc) units or other byproducts were seen. Additionally, a low molecular weight HA was hydrolyzed in 0.1 M DCl for 4 h at 80 degrees C. It was shown that it was primarily the beta-(1-->4)-linkage between GlcNAc and glucuronic acid (GlcA) that was cleaved during hydrolysis at pH < p K a,GlcA. The dependence of the hydrolysis rate constant was further studied as a function of pH between -0.3 and 5. The degradation was found to be random (linear kinetic plots) over the entire pH range studied. Further, the kinetic rate constant was found to depend linearly on pH in the region -0.3 to 3. Above this pH (around the p K a of HA), the kinetic constant decreased more slowly, probably due to either a change in polymer conformation or due to an increased affinity for protons due to the polymer becoming charged as the GlcA units dissociated.  相似文献   

12.
影响海洋鱼产品品质的因素有很多,如鱼体含水量、肉质及传统加工过程中一些原料的用量(如糟制过程中食盐、酒糟等的加入量)、贮藏过程中贮藏条件的改变等,加工和贮藏过程中水分活度、色泽、pH、酸价、过氧化物值、硫化巴比妥酸值、挥发性盐基氮、蛋白质水解程度及微生物等一些指标可以反映海洋鱼产品的品质。本文对影响海洋鱼产品品质的因素及上述指标一些常用的检测方法进行阐述,并简要论述国内外海洋鱼产品的标准化,以期对控制海洋鱼产品品质提供借鉴。  相似文献   

13.
Penicillin acylase from E. coli (EC 3.5.1.11) was found to hydrolyze N-phenylacetylated 1-aminoethylphosphonic acid and its esters. The enzyme preferentially converts the R-form of the substrates: the ratios of the bimolecular rate constants of penicillin acylasecatalyzed hydrolysis of R- and S-forms of 1-(N-phenylacetamino)-ethylphosphonic acid and its dimethyl- and diisopropyl-esters are 58000, 2300, 1800; these derivatives were shown to have the greatest values of the catalytic constants for enzymatic hydrolysis of all known substrates for penicillin acylase: 237, 148 and 134 s-1; the corresponding Km values are 3.7 10(-5), 6.8 10(-4) and 6.2 10(-4) M at pH 7.0. The kinetics of enzymatic hydrolysis of 1-(N-phenylacetamino)-ethylphosphonic acid was investigated up to high degrees of conversion. The inhibition of penicillin acylase by high concentrations of the R-form of the substrate (with substrate inhibition constant of 0.07 M) and competitive inhibition by the reaction product, phenylacetic acid (Ki = 3.5 10(-5) M), was observed.  相似文献   

14.
The preparation of autoradiographs in which the tissue and the emulsion are in permanent register is often complicated by staining after the photographic image has been developed and fixed. While general oversight methods can be satisfactory, controlled, specific staining can be obtained with most basic dyes when the pH is properly regulated. The reactivity of the gelatin is suppressed at a pH of 4 or slightly below whereas nuclei, ergastoplasm, cartilage, mast cells, mucus, etc. stain readily. Basic fuchsin, .05% at pH 3.5-4.0 in dilute (1:10) McLlvaine buffer, is recommended. The final preparation contrasts in color and transparency with the black, opaque silver grains.  相似文献   

15.
A batch reactor was used to investigate the dilute acid hydrolysis reaction of alpha-cellulose and sugar decomposition reactions. Varying the sulfuric acid concentration from 0.07 to 5.0% for reaction temperatures between 180 and 220°C significantly affected glucose yields, which ranged from about 70% to below 10%. Increasing the reaction temperature enhanced this effect. Similar experimental results were obtained for the decomposition of xylose. For sugar decomposition reactions, less than 0.3 g/L of furfural and 5-hydroxymethylfurfural (5-HMF) were produced from glucose and xylose in the absence of sulfuric acid at 190°C and 15 min of reaction time, but adding a small amount of sulfuric acid (0.5%) dramatically increased the decomposition rate and led to the formation of four undesireable products: formic acid, 5-HMF, acetic acid, and furfural. In both hydrolysis and fermentation reactions formic acid, acetic acid, and 5-HMF severely inhibited ethanol fermentation, while furfural had less of an inhibition effect.  相似文献   

16.
A model was developed and evaluated as a tool for predicting the formation of soluble products from staged thermochemical treatment of lignocellulosic materials under acidic conditions typical of autohydrolysis. The model was used to predict the general trend of hemi-cellulose and cellulose hydrolysis between pH 2 and 4 and temperatures of 170-230 degrees C, and results were compared with experimental data. When the model was evaluated for this range of temperatures and pH values, results indicated: (1) a relatively low temperature (175 degrees C) during the first stage allows hydrolysis of the hemi-cellulose polysaccharides without significant mono-saccharide decomposition, (2) subsequent stages at higher temperatures (equal or greater than 200 degrees C) are needed for significant cellulose hydrolysis, but glucose decomposition will also occur, and, (3) a pH in the range of 2-2.5 will enhance polysaccharide hydrolysis while limiting monosaccharide decomposition. The model's predictions, indicating that the formation of biodegradable products could be optimized using Pretreatments at pH 2-2.5 for the pH range evaluated, were confirmed in experiments with white fir as a representative lig nocellulose.  相似文献   

17.
Optically pure (S)-3-hydroxy-gamma-butyrolactone, an important chiral building block in the pharmaceutical industry, was synthesized from L: -malic acid by combining a selective hydrogenation and a lipase-catalyzed hydrolysis. Lipase from Candida rugosa was found to be the most efficient enzyme for the hydrolysis of (S)-beta-benzoyloxy-gamma-butyrolactone. The use of organic solvent-aqueous two-phase system was employed to extract benzoic acid generated from enzymatic hydrolysis of the substrate. Tert-butyl methyl ether as an organic solvent was effective to extract the reaction product, benzoic acid, and stably maintained the enzyme activity of Lipase OF immobilized on polymeric supports Amberlite XAD-7. The immobilization made the recovery of the product simpler and prevented the formation of the emulsion. The pH adjustment was unnecessary with the immobilized Lipase OF. The scale-up of the enzymatic hydrolysis of S-BBL at 1,850-kg scale was carried out without problems to give 728.5 kg of S-HGB at 80% isolated yield. The scale-up results are similar to those of bench scale reactions. Racemic (R,S)-beta-benzoyloxy-gamma-butyrolactone was prepared from D-, L: -malic acid and was found to be hydrolyzed nonselectively by the enzyme.  相似文献   

18.
Mechanism of reaction of 3-hydroxyanthranilic acid with molecular oxygen   总被引:1,自引:0,他引:1  
The autoxidation of the tryptophan metabolite, 3-hydroxyanthranilic acid, at pH 7 gives rise to a p-quinone dimer and cinnabarinic acid. A novel dimer formed by radical-radical coupling of 3-hydroxyanthranilic acid is also produced. Labelling studies have shown that the C-2 oxygen in the p-quinone dimer is derived from molecular oxygen. A product versus time study of this reaction has revealed that, in the absence of catalase, cinnabarinic acid is formed but undergoes decomposition by hydrogen peroxide. At pH 7, in the presence of catalase, both the p-quinone dimer and cinnabarinic acid are formed at approximately the same rate and this rate of formation increases with increasing pH. Inclusion of superoxide dismutase was found to increase the rate of formation of cinnabarinic acid, suggesting that superoxide ions may also cause decomposition of cinnabarinic acid. This was confirmed by treating cinnabarinic acid with superoxide. A mechanism involving a common anthranilyl radical intermediate is proposed to account for the formation of the different oxidation products.  相似文献   

19.
The initiation of haemoglobin synthesis in rabbit reticulocytes   总被引:3,自引:2,他引:1       下载免费PDF全文
1. The incorporation of labelled valine by rabbit reticulocytes into the N-terminal position of nascent haemoglobin was investigated by deaminating the nascent peptides with nitrous acid and isolating labelled alpha-hydroxyisovaleric acid and valine after acid hydrolysis. 2. The amount of radioactivity in alpha-hydroxyisovaleric acid relative to that in valine indicated the presence of 12.3% N-terminal valine having a free amino group. This high value suggests that most if not all nascent peptides contain valine in the N-terminal position. 3. Cell-free preparations containing reticulocyte ribosomes and pH5 enzymes incorporated alpha-hydroxy-[(14)C]isovaleryl-tRNA (where tRNA refers to transfer RNA), which was obtained by deamination of [(14)C]valyl-tRNA from yeast or liver with nitrous acid, into both soluble and nascent protein. 4. When the soluble protein was chromatographed on CM-cellulose, radioactivity was found to be associated with both the alpha-and beta-globin chains. 5. The kinetics of hydrolysis of [(14)C]valine, was also investigated. Most of the material was hydrolysed rapidly at pH10, but a minor component that was relatively stable appeared to be present to the extent of about 10% of the total valyl-tRNA. Valine was, however, the only hydrolysis product detected by paper chromatography. 6. It is concluded that chain initiation in haemoglobin synthesis involves valine as the N-terminal amino acid and that the amino group of nascent protein is probably not substituted.  相似文献   

20.
The effect of pH upon the transpeptidation and hydrolytic reactions of gamma-glutamyltransferase [5-glutamyl)-peptide:amino-acid 5-glutamyltransferase, EC 2.3.2.2) have been investigated. It was found that the enzyme was irreversibly inactivated below pH 7.5 or above pH 9.4. Transpeptidation was markedly pH-dependent, while hydrolysis was pH-independent. The pH optimum for transpeptidation was found to vary for different acceptors. The ascending limb of the pH-optimum curve is attributed to the pK of the alpha-amino group of the acceptor, while the descending limb of the pH-optimum curve is attributed to an ionisable group in the active site of the enzyme. These observations provide much information about the interaction of the enzyme with the acceptor: (1) the true acceptor for gamma-glutamyltransferase is the deprotonated form of the amino acid; (2) glycylglycine has a similar acceptor activity to methionine, its apparent higher activity being due to the low pK of the alpha-amino group; (3) the enzyme is reversibly inactivated at higher pH by the deprotonation of a group in the active site which is involved in both binding of acceptor and catalysis of transpeptidation (this group is not involved in the hydrolysis reaction); (4) at pH 8.5, the normal pH for assay, only 47% of the enzyme is active, while at pH 7.4 gamma-glutamyltransferase is 93% in the active form.  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