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.
The enthalpies of interaction of glucose oxidase at 25°C with a homologous series of n-alkyltrimethylammonium bromides (TABs) at pH 10 and a homologous series of n-alkylsulfates at pH 3.2 have been measured by microcalorimetry. For the n-dodecyl member of each series, DTAB and sodium n-dodecylsulfate (SDS), the binding of the surfactants to glucose oxidase as measured by equilibrium dialysis has been used in combination with the enthalpy data to obtain the Gibbs energy ( ), enthalpy ( ) and entropy ( ) of binding per surfactant molecule as a function of the number of surfactant molecules bound ( ). The thermodynamic parameters for the glucose oxidase interaction with DTAB at pH 10 and SDS at pH 3.2 are very similar and show that the interactions are entropically driven. The observed enthalpies of interaction of glucose oxidase with the homologous n-alkylsulfates have been analysed in terms of the interactions between the anionic surfactant head group and cationic sites on the protein, hydrophobic binding and the thermal contributions arising from protein unfolding. At surfactant concentrations of 0.5 c.m.c., the enthalpy of unfolding of glucose oxidase is estimated to be 3610 ± 560 kJ mol−1.  相似文献   

2.
A series of catalases have been prepared in which a proportion of the carboxyl groups of glutamate and aspartate residues have been amidated with glycinamide. The physical properties of the amidated catalases have been investigated with specific reference to their interaction with sodium n-dodecyl sulphate (SDS). Amidation leads to an increase in SDS binding at pH 6.4. Microcalorimetric measurements show that the exothermic enthalpy of interaction with SDS increases with the extent of amidation in acid solution (pH 3.2–6.4). The increase in exothermicity is compensated by a decrease in entropy since the average Gibbs energy of SDS binding is independent of the extent of amidation. At pH 3.2 where the catalase carboxyl groups are largely un-ionized amidation still increase the exothermicity of the interaction with SDS. It is suggested that at low pH the SDS anion interacts favourably with the resonance stabilized O-protonated form of amidated side chains.  相似文献   

3.
Equine liver glutathione S-transferase has been shown to consist of two identical subunits of apparent Mr 25,500 and a pl of 8.9. Kinetic data at pH 6.5 with 1-chloro-2,4-dinitrobenzene as a substrate suggests a random rapid-equilibrium mechanism, which is supported by inhibition studies using glutathione analogues. S-(p-Bromobenzyl)glutathione and the corresponding N alpha-, CGlu- and CGly-substituted derivatives have been found, at pH 6.5, to be linear competitive inhibitors, with respect to GSH, of glutathione transferase. N-Acetylation of S-(p-bromobenzyl)glutathione decreases binding by 100-fold, whereas N-benzoylation and N-benzyloxycarbonylation abolish binding of the derivative to the enzyme. The latter effect has been attributed to a steric constraint in this region of the enzyme. Amidation of the glycine carboxy group of S-(p-bromobenzyl)glutathione decreases binding by 13-fold, whereas methylation decreases binding by 70-fold, indicating a steric constraint and a possible electrostatic interaction in this region of the enzyme. Amidation of both carboxy groups decreases binding significantly by 802-fold, which agrees with electrostatic interaction of the glutamic acid carboxy group with a group located on the enzyme.  相似文献   

4.
The chemolithoautotroph, Arthrobacter sp.15b oxidizes arsenite to arsenate using a membrane bound arsenite oxidase. The enzyme arsenite oxidase is purified to its homogeneity and identified using MALDI-TOF MS analysis. Upon further characterization, it was observed that the enzyme is a heterodimer showing native molecular mass as ~100 kDa and appeared as two subunits of ~85 kDa LSU and 14 kDa SSU on SDS–PAGE. The V max and K m values of the enzyme was found to be 2.45 μM (AsIII)/min/mg) and 26 μM, respectively. The purified enzyme could withstand wide range of pH and temperature changes. The enzyme, however, gets deactivated in the presence of 1 mM of DEPC suggesting the involvement of histidine at the binding site of the enzyme. The peptide analysis of large sub unit of the enzyme showed close match with the arsenite oxidases of Burkholderia sp. YI019A and arsenite oxidase, Mo-pterin containing subunit of Alcaligenes faecalis. The small subunit, however, differed from other arsenite oxidases and matched only with 2Fe–2S binding protein of Anaplasma phagocytophilum. This indicates that Rieske subunits containing the iron–sulfur clusters present in the large as well as small subunits of the enzyme are integral part of the protein.  相似文献   

5.
Glucose oxidase was immobilized by electropolymerization into films of polyaniline, polyindole, polypyrrole, poly(o-phenylediamine), and polyaniline crosslinked with p-phenylenediamine. The kinetics and the behavior of the entrapped enzyme toward elevated temperature, organic solvent denaturation, and pH were investigated, along with the response of the films to electroactive species such as acetaminophen, ascorbate, cysteine, and uric acid. For most of the films, linearity to glucose extended from 7 to 10 mM. The poly(o-phenylenediamine)/glucose oxidase film gave the best signal/noise ratio and polypyrrole/glucose oxidase film gave the most reproducible current responses. No significant shift of the optimum reaction pH (5.5), except for polypyrrole (5.0), was observed after immobilization of glucose oxidase in the various films. Enzymatic activity decreased rapidly when pH was raised above 7.5. Thermodeactivation studies at 55 degrees , 60 degrees , and 65 degrees C have shown polypyrrole/and poly(o-phenylediamine)/glucose oxidase films to be the most resistant enzymatic films. Poly(o-phenylenediamine) films offered the best protection against glucose oxidase deactivation in hexane, chloroform, ether, THF, and acetonitrile when compared with the other electropolymerized films. All the enzymatic films exhibited permselection toward electroactive species. (c) 1996 John Wiley & Sons, Inc.  相似文献   

6.
The dissociation of beef liver and bacterial (Micrococcus lysodeikticus) catalases by the action of sodium n-dodecyl sulfate (SDS) has been investigated as a function of SDS concetration and time by ultracentrifugation. The rate of dissociation of beef liver catalase is found to be much faster than that for bacterial catalase in 25 mM SDS at pH 7.0. Beef liver catalase is dissociated into its four subunits after 24 h, whereas bacterial catalase is not completely dissociated after 36 days of incubation. The binding of SDS to beef liver catalase obeys a Hill equation with a cooperativity exponent of 2.0 and a binding constant of 440. The initial interaction of SDS with beef liver catalase can be detected by microcalorimetry, whereas the mixing of SDS with bacterial catalase is athermal. Bacterial catalase retains enzymic activity in the presence of SDS, whereas beef liver catalase is completely deactivated at SDS concentrations above 5 mM. Beef liver catalase is more sensitive to acid denaturation than bacterial catalase, and the rate of dissociation for both catalases is sixth-order in proton concentration. Comparison of the amino acid analysis of the two catalases shows that bacterial catalase has a smaller number of lysyl residues and a larger number of glutamyl residues than beef liver catalase. Taken together these structural differences could lead to a reduced affinity of bacterial catalase for the binding of SDS as observed.  相似文献   

7.
Glucose oxidase from Aspergillus niger was purified to homogeneity by hydrophobic interaction and ion-exchange chromatography. Approx. 95% of the carbohydrate moiety was cleaved from the protein by incubation of glucose oxidase with endoglycosidase H and alpha-mannosidase. Cleavage of the carbohydrate moiety effected a 24-30% decrease in the molecular weight and a reduction in the number of isoforms of glucose oxidase. No significant changes were observed in the circular dichroism spectra of the deglycosylated enzyme. Other properties, such as thermal stability, pH and temperature optima of glucose oxidase activity and substrate specificity were not affected. However, removal of the carbohydrate moiety marginally affected the kinetics of glucose oxidation and stability at low pH. From these results it appears that the carbohydrate chain of glucose oxidase does not contribute significantly to the structure, stability and activity of glucose oxidase.  相似文献   

8.
Glucose 2-oxidase (EC 1.1.3.10) from Coriolus versicolor and Phanerochaete chrysosporium and glucose 1-oxidase (EC 1.1.3.4) from Aspergillus niger bound to a CU(II)-IDA column in the pH range of 6–8. However, glucose 1-oxidase from Penicillium amagasakiense bound only partially to a CU(II)-IDA column at pH 8.0. Metal chelates containing either Ni(II) or Zn(II) were useful in the adsorption of glucose 2-oxidase from Phanerochaete chrysosporium. The binding of glucose 2-oxidase from P. chrysosporium to Ni(II) and Zn(II)-IDA agarose columns increases as a function of pH of the buffer system. The adsorption of glucose oxidases on metal(II)-IDA chelates was due to the available histidine residues on enzyme molecules since the addition of imidazole in the buffer system abolished the binding of glucose oxidases to these columns. Both glucose oxidases from C.versicolor, P. chrysosporium and A. niger were purified in one step by immobilized metal affinity chromatography on metal(II)-IDA agarose columns with a recovery of enzyme activity in the range of 80–91%. Purified preparations of glucose oxidases from fungal strains were apparently homogeneous on native PAGE and SDS-PAGE. Immobilized metal affinity chromatography was used to separate glucose 1-oxidase from the 2-oxidase on metal(II)-IDA agarose columns which was confirmed by analysis of the reaction products by HPLC. The different chromatographic behaviour of glucose oxidases on metal(II)-IDA chelates is apparently due to the number and spatial distribution of available histidine residues on these enzyme molecules. Received 12 May 1998/ Accepted in revised form 29 July 1998  相似文献   

9.
1. From the standpoint of monomer-dimer equilibrium of hog kidney D-amino acid oxidase [EC 1.4.3.3] and the interaction between the enzyme and small molecules, the effect of pH on the binding of p-aminobenzoate to the monomer and dimer of the enzyme was studied by kinetic methods and spectrophotometric titration. 2. The maximum binding number of p-aminobenzoate to the dimer is two molecules, and there is no interaction between the two active sites of the dimer (i.e., no cooperativity) over the range of pH from 6.5 to 10. 3. The affinity of the dimer for p-aminobenzoate is several times higher than that of the monomer at pH 6.5-10, and consequently p-aminobenzoate induces dimerization in the equilibrium state of D-amino acid oxidase. The interaction energy of two subunits of the dimer is stabilized by the binding of p-aminobenzoate by 1-2 kcal/mole over the pH range studied. 4. The binding sites of the quasi-substrate, p-aminobenzoate, in the dimer and the intersubunit binding site of the dimer are clearly different, because p-aminobenzoate induces dimerization of the enzyme. 5. The pK values of ionizing groups in the free monomer and the free dimer which participate in the binding of the competitive inhibitor, p-aminobenzoate, are approximately the same, 8.7, as determined from the pH dependence of the affinity of the inhibitor for the enzyme. Furthermore, no pK for the enzyme-inhibitor complex in the pH range 6.5-10 was observed. 6. There is no interaction between the two ionizing groups of the dimer during protonation-deprotonation, because a theoretical equation involving no cooperativity between the two ionizing groups in the dimer explains the results well.  相似文献   

10.
A factor in medium conditioned by mouse tumor cells was shown previously to suppress the capacity of mouse peritoneal macrophages to undergo a respiratory burst and to kill protozoal pathogens (macrophage deactivation factor, MDF). Recently, pure transforming growth factor-beta (TGF-beta) proved to be a potent macrophage deactivator as well. Two lines of evidence suggest that MDF is not identical with TGF-beta. First, rabbit anti-TGF-beta IgG neutralized the respiratory burst-suppressing activity of TGF-beta without affecting the bioactivity of MDF, even when the latter was treated with acid to activate potentially latent TGF-beta. Second, in contrast to MDF, which decreases the affinity of the NADPH oxidase for NADPH, permeabilized macrophages that had been deactivated with TGF-beta displayed the same Km and Vmax of the oxidase as activated macrophages. As with MDF, TGF-beta had no effect on two other potential control points over the secretion of respiratory burst products, namely, hydrogen peroxide catabolism or glucose uptake. Finally, neither MDF nor TGF-beta affected the extent or affinity of binding of phorbol diesters to macrophages, the activity or cofactor requirements for protein kinase C, or the ability of protein kinase C to translocate quantitatively from cytosol to membrane fractions in response to phorbol diesters. Thus, 1) MDF is not identical with TGF-beta, and 2) in contrast to the activation or deactivation of macrophages by numerous other agents, TGF-beta regulates macrophage respiratory burst capacity at a level other than the apparent affinity of the oxidase for its substrate.  相似文献   

11.
A method for isolating extracellular glucose oxidase from the fungus Penicillium funiculosum 46.1, using ultrafiltration membranes, was developed. Two samples of the enzyme with a specific activity of 914-956 IU were obtained. The enzyme exhibited a high catalytic activity at pH above 6.0. The effective rate constant of glucose oxidase inactivation at pH 2.6 and 16 degrees C was 2.74 x 10(-6) s-1. This constant decreased significantly as pH of the medium increased (4.0-10.0). The temperature optimum for glucose oxidase-catalyzed beta-D-glucose oxidation was in the range 30-65 degrees C. At temperatures below 30 degrees C, the activation energy for beta-D-glucose oxidation was 6.42 kcal/mol; at higher temperatures, this parameter was equal to 0.61 kcal/mol. Kinetic parameters of glucose oxidase-catalyzed delta-D-glucose oxidation depended on the initial concentration of the enzyme in the solution. Glucose oxidase also catalyzed the oxidation of 2-deoxy-D-glucose, maltose, and galactose.  相似文献   

12.
The respiratory chain of the thermohalophilic bacterium Rhodothermus marinus contains an oxygen reductase, which uses HiPIP (high potential iron-sulfur protein) as an electron donor. The structural genes encoding the four subunits of this HiPIP:oxygen oxidoreductase were cloned and sequenced. The genes for subunits II, I, III, and IV (named rcoxA to rcoxD) are found in this order and seemed to be organized in an operon of at least five genes with a terminator structure a few nucleotides downstream of rcoxD. Examination of the amino acid sequence of the Rcox subunits shows that the subunits of the R. marinus enzyme have homology to the corresponding subunits of oxidases belonging to the superfamily of heme-copper oxidases. RcoxB has the conserved histidines involved in binding the binuclear center and the low-spin heme. All of the residues proposed to be involved in proton transfer channels are conserved, with the exception of the key glutamate residue of the D-channel (E(278), Paracoccus denitrificans numbering). Analysis of the homology-derived structural model of subunit I shows that the phenol group of a tyrosine (Y) residue and the hydroxyl group of the following serine (S) may functionally substitute the glutamate carboxyl in proton transfer. RcoxA has an additional sequence for heme C binding, after the Cu(A) domain, that is characteristic of caa(3) oxidases belonging to the superfamily. Homology modeling of the structure of this cytochrome domain of subunit II shows no marked electrostatic character, especially around the heme edge region, suggesting that the interaction with a redox partner is not of an electrostatic nature. This observation is analyzed in relation to the electron donor for this caa(3) oxidase, the HiPIP. In conclusion, it is shown that an oxidase, which uses an iron-sulfur protein as an electron donor, is structurally related to the caa(3) class of heme-copper cytochrome c oxidases. The data are discussed in the framework of the evolution of oxidases within the superfamily of heme-copper oxidases.  相似文献   

13.
The cytochrome o complex is one of two ubiquinol oxidases in the aerobic respiratory system of Escherichia coli. This enzyme catalyzes the two-electron oxidation of ubiquinol-8 which is located in the cytoplasmic membrane, and the four-electron reduction of molecular oxygen to water. The purified oxidase contains at least four subunits by sodium dodecyl sulfate-polyacrylamide gel electrophoresis analysis and has been shown to couple electron flux to the generation of a proton motive force across the membrane. In this paper, the DNA sequence of the cyo operon, containing the structural genes for the oxidase, is reported. This operon is shown to encode five open reading frames, cyoABCDE. The gene products of three of these, cyoA, cyoB, and cyoC, are clearly related to subunits II, I, and III, respectively, of the eukaryotic and prokaryotic aa3-type cytochrome c oxidases. This family of cytochrome c oxidases contain heme a and copper as prosthetic groups, whereas the E. coli enzyme contains heme b (protoheme IX) and copper. The most striking sequence similarities relate the large subunits (I) of both the E. coli quinol oxidase and the cytochrome c oxidases. It is likely that the sequence similarities reflect a common molecular architecture of the two heme binding sites and of a copper binding site in these enzymes. In addition, the cyoE open reading frame is closely related to a gene denoted ORF1 from Paracoccus dentrificans which is located in between the genes encoding subunits II and III of the cytochrome c oxidase of this organism. The function of the ORF1 gene product is not known. These sequence relationships define a superfamily of membrane-bound respiratory oxidases which share structural features but which have different functions. The E. coli cytochrome o complex oxidizes ubiquinol but has no ability to catalyze the oxidation of reduced cytochrome c. Nevertheless, it is clear that the E. coli oxidase and the aa3-type cytochrome c oxidases must have very similar structures, at least in the vicinity of the catalytic centers, and they are very likely to have similar mechanisms for bioenergetic coupling (proton pumping).  相似文献   

14.
An immobilization method using chitosan prepared from chitin as an insoluble carrier was investigated. Glucose isomerase, urease, glucamylase, trypsin and glucose oxidase were attached to chitosan by the aid of water soluble carbodiimide. Their activity yields were as follows; glucose isomerase 32%, urease 44%, glucamylase 8%, trypsin 10%, glucose oxidase 37%.

Immobilized glucose isomerase showed no significant changes in optimal temperature and heat stability. But pH optimum of reaction and pH stability range were somewhat lowered. The inhibitory effects of bivalent metal ions were considerably reduced by immobilization and similar tendency was observed for buffer reagents such as Tris or veronal. Immobilized glucose isomerase was inhibited by 8 m urea or 6 m guanidine hydrochloride in nearly the same way as free enzyme. With SDS, cysteine or mercaptoethanol free glucose isomerase was scarcely affected by these reagents, while immobilized enzyme considerably suffered to a loss of its activity.  相似文献   

15.
Glucose oxidase from Aspergillus niger, hyaluronidase from Streptomyces hyalurolyticus, and cholesterol oxidase and cholesterol esterase from Pseudomonas fluorescens were effectively adsorbed on an Amberlite CG-50 column, when the cell-free cultured medium or the cultured medium with cell extract and without cell debris was applied without desalting but at pH less than or equal to 4.5. At the acidic pH, all the ion-exchange groups (-COOH) exist in the protonated form; the adsorption is not due to electrostatic attraction, but to hydrophobic interaction. The enzymes thus adsorbed were effectively eluted by increasing pH, at which the ion-exchange groups became dissociated. This type of adsorption-elution is called hydrophobic-ionic chromatography. By a single run of chromatography, glucose oxidase, hyaluronidase, cholesterol oxidase, and cholesterol esterase were purified 30-fold, 12-fold, 45-fold, and 20-fold with yields of 82%, 83%, 80%, and 90%, respectively. This indicates that hydrophobic-ionic chromatography on an Amberlite CG-50 column is effective for the purification of various enzymes, provided that they are stable at the acidic pH.  相似文献   

16.
Q Su  J P Klinman 《Biochemistry》1999,38(26):8572-8581
Glucose oxidase catalyzes the oxidation of glucose by molecular dioxygen, forming gluconolactone and hydrogen peroxide. A series of probes have been applied to investigate the activation of dioxygen in the oxidative half-reaction, including pH dependence, viscosity effects, 18O isotope effects, and solvent isotope effects on the kinetic parameter Vmax/Km(O2). The pH profile of Vmax/Km(O2) exhibits a pKa of 7.9 +/- 0.1, with the protonated enzyme form more reactive by 2 orders of magnitude. The effect of viscosogen on Vmax/Km(O2) reveals the surprising fact that the faster reaction at low pH (1.6 x 10(6) M-1 s-1) is actually less diffusion-controlled than the slow reaction at high pH (1.4 x 10(4) M-1 s-1); dioxygen reduction is almost fully diffusion-controlled at pH 9.8, while the extent of diffusion control decreases to 88% at pH 9.0 and 32% at pH 5.0, suggesting a transition of the first irreversible step from dioxygen binding at high pH to a later step at low pH. The puzzle is resolved by 18O isotope effects. 18(Vmax/Km) has been determined to be 1.028 +/- 0.002 at pH 5.0 and 1.027 +/- 0.001 at pH 9.0, indicating that a significant O-O bond order decrease accompanies the steps from dioxygen binding up to the first irreversible step at either pH. The results at high pH lead to an unequivocal mechanism; the rate-limiting step in Vmax/Km(O2) for the deprotonated enzyme is the first electron transfer from the reduced flavin to dioxygen, and this step accompanies binding of molecular dioxygen to the active site. In combination with the published structural data, a model is presented in which a protonated active site histidine at low pH accelerates the second-order rate constant for one electron transfer to dioxygen through electrostatic stabilization of the superoxide anion intermediate. Consistent with the proposed mechanisms for both high and low pH, solvent isotope effects indicate that proton transfer steps occur after the rate-limiting step(s). Kinetic simulations show that the model that is presented, although apparently in conflict with previous models for glucose oxidase, is in good agreement with previously published kinetic data for glucose oxidase. A role for electrostatic stabilization of the superoxide anion intermediate, as a general catalytic strategy in dioxygen-utilizing enzymes, is discussed.  相似文献   

17.
The deactivation of the bovine G-protein-coupled receptor, rhodopsin, is a two-step process consisting of the phosphorylation of specific serine and threonine residues in the cytoplasmic tail of rhodopsin by rhodopsin kinase. Subsequent binding of the regulatory protein arrestin follows this phosphorylation. Previous results find that at least three phosphorylatable sites on the rhodopsin tail (T340) and at least two of the S338, S334, or S343 sites are needed for complete arrestin-mediated deactivation. Thus, to elucidate the details of the interaction between rhodopsin with arrestin, we have employed both a computational and an in vitro experimental approach. In this work, we first simulated the interaction of the carboxy tail of rhodopsin with arrestin using a Monte Carlo simulated annealing method. Since at this time phosphorylation of specific serines and threonines is not possible in our simulations, we substitute either aspartic or glutamic acid residues for the negatively charged phosphorylated residues required for binding. A total of 17 simulations were performed and analysis of this shows specific charge-charge interactions of the carboxy tail of rhodopsin with arrestin. We then confirmed these computational results with assays of comparable constructed rhodopsin mutations using our in vitro assay. This dual computational/experimental approach indicates that sites S334, S338, and T340 in rhodopsin and K14 and K15 on arrestin are indeed important in the interaction of rhodopsin with arrestin, with a possible weaker S343 (rhodopsin)/K15 (arrestin) interaction.  相似文献   

18.
Summary Nylon tube was directly activated by alkaline glutaraldehyde solution. PEI was utilised as a spacer molecule. Glucose oxidase was immobilised to the nylon tube after reactivating the spacer molecules with glutaraldehyde. On immobilising glucose oxidase there was more protein binding and higher immobilised enzyme activity when compared to immobilised enzyme tube activated by triethyloxonium salt. The optimal condition for direct glutaraldehyde activation of nylon was incubation with 18.5% (w/v) glutaraldehyde in 0.12M borate pH 9.0 for 15 min at 90 °.  相似文献   

19.
Human sulphamate sulphohydrolase was purified at least 20,000-fold to homogeneity from liver with a three-step four-column procedure, which consisted of a concanavalin A-Sepharose/Blue A agarose coupled step, and Bio-Gel HT step and then a CM-Sepharose step. The procedure was also used to purify enzyme from kidney and placenta. The subunit Mr of liver, kidney and placenta sulphamate sulphohydrolase was assessed to be 56,000 by using SDS/polacrylamide-gel electrophoresis. The native protein Mr of enzyme from all three tissue sources was assessed by gel-permeation chromatography to be approx. 120,000 on Sephacryl S-300 and 100,000 on Fractogel TSK. It is probable that the native enzyme results from dimerization of subunits. Kinetic parameters (km and kcat.) of human liver sulphamate sulphohydrolase were determined with a variety of substrates matching structural aspects of the physiological substrates in vivo, namely heparin and heparan sulphate. More structurally complex substrates, in which several aspects of the aglycone structure of the natural substrate were maintained, are turned over up to 372000 times faster than the monosaccharide substrate 2-sulphaminoglucosamine. Aglycone structures that influence substrate binding and/or enzyme activity were penultimate-residue C-6 carboxy and C-2 sulphate ester groups and a post-penultimate 2-sulphaminoglucosamine residue. The C-4 hydroxy group of the 2-sulphaminoglucosamine under enzymic attack is involved in binding of substrate to enzyme. The presence of C-6 sulphate ester on the non-reducing end 2-sulphaminoglucosamine stimulates sulphamate bond hydrolysis and substrate affinity if the adjacent monosaccharide residue is idose or 2-sulphoidose, but strongly inhibits hydrolysis if the adjacent monosaccharide residue is iduronic acid. Sulphamate sulphohydrolase is an exoenzyme, since activity toward internal sulphamate bonds was not detected. The effect of incubation pH on enzyme activity towards the variety of substrates evaluated was complex and dependent on substrate aglycone structure. The presence of aglycone C-2 sulphate ester and aglycone C-6 carboxy groups and C-6 sulphate ester groups on the 2-sulphaminoglucosamine residue under attack considerably affect the pH response. Structurally complex substrates had two pH optima. Incubation temperature and buffer ionic strength markedly influenced pH optima and enzyme activity. Cu2+ and SO4(2-)ions are potent inhibitors of enzyme activity.  相似文献   

20.
Glucose oxidase (GOD) was covalently immobilized on amorphous AlPO4 as well as on an AlPO4/clay mineral Sepiolite system. Immobilization of the enzyme was carried out through the -amino group of lysine residues through an aromatic Schiff's-base. Activation of the support was obtained after reaction of appropriate molecules with support surface –OH groups. The enzymatic activities of native, and different immobilized GOD systems and filtrates, were followed by the amount of liberated -gluconic acid obtained in the enzymatic β- -glucose oxidation with the aid of an automatic titrator. The kinetic properties of native and immobilized GOD were obtained for glucose concentrations in the range of physiological conditions and at different working conditions such as reaction temperature, reaction pH, and enzyme concentration.

The binding percentage of enzymes was in the 50–80% range, with residual and specific activities in the 65–80% and 90–150% ranges, respectively. No change in the pH optimum and only slight changes in the Vmax and KM kinetic parameters with respect to native GOD were observed, so that not only was little deactivation of enzyme obtained throughout the immobilization process but also that the stability of the covalently bound enzyme in the two supports appeared to have increased with respect to the soluble enzyme. GOD immobilization also increased its efficiency and operational stability in repeated uses on increasing the amount of immobilized enzyme.  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