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.
We investigated water/organic solvent sorption and residual enzyme activity to simultaneously monitor preferential solvation/hydration of protein macromolecules in the entire range of water content at 25°C. We applied this approach to estimate protein destabilization/stabilization due to the preferential interactions of bovine pancreatic α‐chymotrypsin with water‐acetone (moderate‐strength H‐bond acceptor) and water‐DMSO (strong H‐bond acceptor) mixtures. There are three concentration regimes for the dried α‐chymotrypsin. α‐Chymotrypsin is preferentially hydrated at high water content. The residual enzyme activity values are close to 100%. At intermediate water content, the dehydrated α‐chymotrypsin has a higher affinity for acetone/DMSO than for water. Residual enzyme activity is minimal in this concentration range. The acetone/DMSO molecules are preferentially excluded from the protein surface at the lowest water content, resulting in preferential hydration. The residual catalytic activity in the water‐poor acetone is ~80%, compared with that observed after incubation in pure water. This effect is very small for the water‐poor DMSO. Two different schemes are operative for the hydrated enzyme. At high and intermediate water content, α‐chymotrypsin exhibits preferential hydration. However, at intermediate water content, in contrast to the dried enzyme, the initially hydrated α‐chymotrypsin possesses increased preferential hydration parameters. At low water content, no residual enzyme activity was observed. Preferential binding of DMSO/acetone to α‐chymotrypsin was detected. Our data clearly demonstrate that the hydrogen bond accepting ability of organic solvents and the protein hydration level constitute key factors in determining the stability of protein–water–organic solvent systems.  相似文献   

2.
Sodium alginate, activated by periodate oxidation, was covalently linked to porcine pancreatic α‐amylase via reductive alkylation with NaBH4. The enzyme‐polymer conjugate, purified by gel filtration on Fractogel EMD BioSEC (S), retained about 50% of the native specific amylolytic activity. The sugar content was estimated to be 712 mol of monosaccharides per mol of enzyme protein. An average of 11 amino groups out of 21 groups from α‐amylase were modified with the polysaccharide. The functional stability was improved for α‐amylase after conjugation with sodium alginate. In comparison with the native enzyme, the thermostability of α‐amylase was increased by this modification. In addition, the stability in the range of pH 5.0–11.0 was improved for the modified enzyme. The conjugate was also more resistant to denaturation by 0.3% sodium dodecylsulphate, retaining about 10% of its initial activity after 120 min of incubation. The formation of stabilizing salt bridges in the protein surface of the α‐amylase‐polysaccharide complex was confirmed by FT‐IR spectrometry. Attending to the results obtained, we conclude that the covalent attachment of the anionic polysac‐charide sodium alginate to the enzymes might be a useful and non‐expensive method for improving the stabilization of these biocatalysts under various denaturing conditions.  相似文献   

3.
Two different α‐glucosidase‐producing thermophilic E134 strains were isolated from a hot spring in Kozakli, Turkey. Based on the phenotypic, phylogenetic and chemotaxonomic evidence, the strain was proposed to be a species of G. toebii. Its thermostable exo‐α‐1,4‐glucosidases also were characterized and compared, which were purified from the intracellular and extracellular fractions with estimated molecular weights of 65 and 45 kDa. The intracellular and extracellular α‐glucosidases showed optimal activity at 65 °C, pH 7·0, and at 70 °C, pH 6·8, with 3·65 and 0·83 Km values for the pNPG substrate, respectively. Both enzymes remained active over temperature and pH ranges of 35–70 °C and 4·5–11·0. They retained 82 and 84% of their activities when incubated at 60 °C for 5 h. Their relative activities were 45–75% and 45–60% at pH 4·5 and 11·0 values for 15 h at 35 °C. They could hydrolyse the α‐1,3 and α‐1,4 bonds on substrates in addition to a high transglycosylation activity, although the intracellular enzyme had more affinity to the substrates both in hydrolysis and transglycosylation reactions. Furthermore, although sodium dodecyl sulfate behaved as an activator for both of them at 60 °C, urea and ethanol only increased the activity of the extracellular α‐glucosidase. By this study, G. toebii E134 strain was introduced, which might have a potential in biotechnological processes when the conformational stability of its enzymes to heat, pH and denaturants were considered. Copyright © 2011 John Wiley & Sons, Ltd.  相似文献   

4.
Four new carbocyanines containing symmetric and asymmetric heterocyclic moieties and N‐carboxyalkyl groups have been synthesized and characterized. The binding mechanism established between these cyanines and several proteins was evaluated using saturation transfer difference (STD) NMR. The results obtained for the different dyes revealed a specific interaction to the standard proteins lysozyme, α‐chymotrypsin, ribonuclease (RNase), bovine serum albumin (BSA), and gamma globulin. For instance, the two un‐substituted symmetrical dyes (cyanines 1 and 3) interacted preferentially through its benzopyrrole and dibenzopyrrole units with lysozyme, α‐chymotrypsin, and RNase, whereas the symmetric disulfocyanine dye (cyanine 2) bound BSA and gamma globulin through its carboxyalkyl chains. On the other hand, the asymmetric dye (cyanine 4) interacts with lysozyme and α‐chymotrypsin through benzothiazole moiety and with RNase through dibenzopyrrole unit. Thus, STD‐NMR technique was successfully used to screen cyanine–protein interactions and determine potential binding sites of the cyanines for posterior use as ligands in affinity chromatography. Copyright © 2014 John Wiley & Sons, Ltd.  相似文献   

5.
A series of novel α‐(diphenylphosphoryl)‐ and α‐(diphenylphosphorothioyl)cycloalkanone oximes have been synthesized in search for novel bioactive molecules. Their structures were characterized by various spectroscopic methods including IR, NMR (1H, 31P, 13C), mass spectrometry and single crystal X‐ray diffraction. The newly synthesized phosphorus‐containing oximes were screened for their in vitro antimicrobial activity against Gram‐positive bacteria (Staphylococcus aureus and Bacillus subtilis), Gram‐negative bacteria (Escherichia coli and Salmonella typhimurium) and fungal strains (Candida albicans and Candida glabrata). The biological assays showed that all the studied compounds exhibited high antibacterial and antifungal activities at only 0.1–2.1 μg/mL. In silico molecular docking studies in FabH enzyme active site were performed in order to predict the possible interaction modes and binding energies of the drug candidates at the molecular level.  相似文献   

6.
Cleavage of the ether bond of chlorophenoxyalkanoate herbicides is catalyzed by an α‐ketoglutarate‐linked dioxygenase (TfdA). In this step, α‐ketoglutarate is decarboxylated to succinate and must be regenerated for continual substrate cleavage. Limitations in herbicide degradation are to be expected in the case of a shortage of α‐ketoglutarate. Such a situation was simulated and studied with Delftia (formerly Comamonas) acidovorans MC1 and Rhodoferax sp. P230, which constitutively express etherolytic dioxygenase activity by excreting 2,4‐dichlorophenol (DCP) as a dead‐end product. The results showed that 2,4‐dichlorophenoxyacetate (2,4‐D) could hardly be cleaved under these conditions which is attributed to the inability to regenerate α‐ketoglutarate from the cleavage products, i.e. succinate and glyoxylate [1 ]. With pyruvate, in contrast, liberated as the oxidized alkanoic acid from the cleavage of (RS )‐2‐(2,4‐dichlorophenoxy)propionate (2,4‐DP), the regeneration of α ‐ketoglutarate seems to be guaranteed from succinate as resulted from the utilization of 2,4‐DP to a considerable amount under these conditions. The extent was limited, however, which was apparently caused by the accumulation of DCP. Continual cleavage of 2,4‐DP could be demonstrated in the presence of Ochrobactrum sp. K2‐14, which functions as a DCP‐consuming strain. Addition of extra metabolites, i.e. α‐ketoglutarate or other readily metabolizable substrates, improved the cleavage of the herbicides. This was most pronounced with 2,4‐D that was found now to be also utilized to a considerable extent. Conversely, the cleavage of the herbicides (2,4‐DP) was reduced and ultimately ceased with cells depleted by starvation of the pool of metabolites. Again, this deficit could be restored by adding α‐ketoglutarate. The limitations in utilizing phenoxyalkanoate herbicides are discussed in terms of pseudo‐recalcitrance owing to deficits in metabolites (α‐ketoglutarate) rather than enzyme activity (TfdA).  相似文献   

7.
The activity of a lipase from a newly isolated Pseudomonas sp. was investigated in the presence of organic solvents and imidazolium chloride‐based ionic liquids (IL) such as BMIM[Cl] and HMIM[Cl]. The lipase activity in the presence of IL was higher compared to that in common organic solvents such as methanol and 2‐propanol. A possible explanation for the enzyme activation might be the structural changes induced in the protein in organic systems. Since IL quench the intensity of fluorescence emission, it was not possible to investigate the major factor that influences the enzyme behavior in these new organic salts. Furthermore, the enzyme exhibited excellent activity in buffer mixtures containing both organic solvent and IL. The stability of the lipase at 50°C was considerably increased in the presence of 20% BMIM[Cl] compared with the untreated lipase in aqueous medium. The light scattering method clearly showed that prevention of aggregation could be the reason for thermal stabilization at 50°C in reactions containing IL. Kinetic analysis of the enzyme in the presence of different concentrations of IL showed that the Km value increased from 0.45 mM in aqueous buffer to 2.4 mM in 50% v/v BMIM[Cl]/buffer. The increase in Km indicates that IL can significantly reduce the binding affinity of the substrate to the enzyme. Also, a linear correlation was observed between the BMIM[Cl] concentration and Vmax of the enzyme. As the concentration of BMIM[Cl] increased from 10 to 50% v/v, the Vmax value increased from 1.8 to 46 μM/min.  相似文献   

8.
The ‘lipid‐coated ice‐droplet hydration method’ was applied for the preparation of milliliter volumes of a suspension of giant phospholipid vesicles containing in the inner aqueous vesicle pool in high yield either calcein, α‐chymotrypsin, fluorescently labeled bovine serum albumin or dextran (FITC‐BSA and FITC‐dextran; FITC=fluorescein isothiocyanate). The vesicles had an average diameter of ca. 7–11 μm and contained 20–50% of the desired molecules to be entrapped, the entrapment yield being dependent on the chemical structure of the entrapped molecules and on the details of the vesicle‐formation procedure. The ‘lipid‐coated ice droplet hydration method’ is a multistep process, based on i) the initial formation of a monodisperse water‐in‐oil emulsion by microchannel emulsification, followed by ii) emulsion droplet freezing, and iii) surfactant and oil removal, and replacement with bilayer‐forming lipids and an aqueous solution. If one aims at applying the method for the entrapment of enzymes, retention of catalytic activity is important to consider. With α‐chymotrypsin as first model enzyme to be used with the method, it was shown that high retention of enzymatic activity is possible, and that the entrapped enzyme molecules were able to catalyze the hydrolysis of a membrane‐permeable substrate which was added to the vesicles after their formation. Furthermore, one of the critical steps of the method that leads to significant release of the molecules from the water droplets was investigated and optimized by using calcein as fluorescent probe.  相似文献   

9.
A simple, rapid and reproducible procedure for the identification of extracellular cucumber (Cucumis sativus L.) α‐galactosidase is described using callus cultures of seedlings from the tested plant, hairy roots of 2‐day‐old seedlings of cucumber germinating on agar plates as well as cell suspension cultures derived from callus cultures. For the determination of the intracellular and extracellular activities of α‐galactosidase, 6‐bromo‐2‐naphthyl‐αD‐galactopyranoside and p‐nitrophenyl‐αD‐galactopyranoside, respectively, were used as synthetic substrates. The extracellular α‐galactosidase activity was identified by evaluating the dye‐zones in agar medium. The enzyme from cucumber callus cultures and seedling roots, cultivated on agar plates supplemented with 6‐bromo‐2‐naphthyl‐αD‐galactopyranoside, hydrolyzed this substrate releasing 6‐bromo‐2‐naphthol. By simultaneous coupling with hexazonium p‐rosaniline the corresponding azodye was formed. Thus, the extracellular enzyme was detected by the presence of reddish‐brown zones on the agar plates around the plant material. The parallel extracellular and intracellular activities were determined in cell suspension cultures derived from callus cultures. The results show a 44.6% intracellular and 55.4% extracellular distribution of α‐galactosidase activity. The described agar plate method enables a rapid, simple and specific detection of plant producers of extracellular α‐galactosidase.  相似文献   

10.
Protonation/deprotonation equilibria are frequently linked to binding processes involving proteins. The presence of these thermodynamically linked equilibria affects the observable thermodynamic parameters of the interaction (K obs, ΔH obs0). In order to try and elucidate the energetic factors that govern these binding processes, a complete thermodynamic characterisation of each intrinsic equilibrium linked to the complexation event is needed and should furthermore be correlated to structural information. We present here a detailed study, using NMR and ITC, of the interaction between α-chymotrypsin and one of its competitive inhibitors, proflavin. By performing proflavin titrations of the enzyme, at different pH values, we were able to highlight by NMR the effect of the complexation of the inhibitor on the ionisable residues of the catalytic triad of the enzyme. Using ITC we determined the intrinsic thermodynamic parameters of the different equilibria linked to the binding process. The possible driving forces of the interaction between α-chymotrypsin and proflavin are discussed in the light of the experimental data and on the basis of a model of the complex. This study emphasises the complementarities between ITC and NMR for the study of binding processes involving protonation/deprotonation equilibria. Electronic supplementary material The online version of this article (doi:) contains supplementary material, which is available to authorized users.  相似文献   

11.
The use of quail meat and eggs has made this animal important in recent years, with its low cost and high yields. Glutathione S‐transferases (GST, E.C.2.5.1.18) are an important enzyme family, which play a critical role in detoxification system. In our study, GST was purified from quail liver tissue with 47.88‐fold purification and 12.33% recovery by glutathione agarose affinity chromatography. The purity of enzyme was checked by SDS‐PAGE method and showed a single band. In addition, inhibition effects of (3aR,4S,7R,7aS)‐2‐(4‐((E)‐3‐(aryl)acryloyl)phenyl)‐3a,4,7,7a‐tetrahydro‐1H‐4,7methanoisoindole‐1,3(2H)‐dion derivatives ( 1a–g ) were investigated on the enzyme activity. The inhibition parameters (IC50 and Ki values) were calculated for these compounds. IC50 values of these derivatives ( 1a–e ) were found as 23.00, 15.75, 115.50, 10.00, and 28.75 μM, respectively. Ki values of these derivatives ( 1a–e ) were calculated in the range of 3.04 ± 0.50 to 131.50 ± 32.50 μM. However, for f and g compounds, the inhibition effects on the enzyme were not found.  相似文献   

12.
In the presence of hemin and under appropriate conditions, some modalities of G‐quadruplexes can form a peroxidase‐like DNAzyme that has been widely used in biology. Structure? function studies on the DNAzyme revealed that its catalytic ability may be dependent on the unimolecular parallel G‐quadruplex. In this report, we present the preliminary investigation on the relationship between the structure and function of DNAzymes through a terminal oligo modification in G‐quadruplex sequences by adding different lengths of oligo‐dT to the 3′‐ or 5′‐end of the aptamers. The results suggested that adding dTn to the 5′‐end of the DNA sequence of the enzyme improved the ability of hemin to bind with DNA, but the addition of dTn to the 3′‐end decreased the binding ability of hemin for DNA. The increased stability of the assembled DNAzyme would lead to more favorable binding between the enzyme and substrate (H2O2), facilitating higher peroxidase activity; on the contrary, with lower stability of the DNAzyme complex, we observed reduced peroxidase activity.  相似文献   

13.
7‐Carboxy‐7‐deazaguanine synthase, QueE, catalyzes the radical mediated ring contraction of 6‐carboxy‐5,6,7,8‐tetrahydropterin, forming the characteristic pyrrolopyrimidine core of all 7‐deazaguanine natural products. QueE is a member of the S‐adenosyl‐L‐methionine (AdoMet) radical enzyme superfamily, which harnesses the reactivity of radical intermediates to perform challenging chemical reactions. Members of the AdoMet radical enzyme superfamily utilize a canonical binding motif, a CX3CX?C motif, to bind a [4Fe‐4S] cluster, and a partial (β/α)6 TIM barrel fold for the arrangement of AdoMet and substrates for catalysis. Although variations to both the cluster‐binding motif and the core fold have been observed, visualization of drastic variations in the structure of QueE from Burkholderia multivorans called into question whether a re‐haul of the defining characteristics of this superfamily was in order. Surprisingly, the structure of QueE from Bacillus subtilis revealed an architecture more reminiscent of the classical AdoMet radical enzyme. With these two QueE structures revealing varying degrees of alterations to the classical AdoMet fold, a new question arises: what is the purpose of these alterations? Here, we present the structure of a third QueE enzyme from Escherichia coli, which establishes the middle range of the spectrum of variation observed in these homologs. With these three homologs, we compare and contrast the structural architecture and make hypotheses about the role of these structural variations in binding and recognizing the biological reductant, flavodoxin. Broader impact statement: We know more about how enzymes are tailored for catalytic activity than about how enzymes are tailored to react with a physiological reductant. Here, we consider structural differences between three 7‐carboxy‐7‐deazaguanine synthases and how these differences may be related to the interaction between these enzymes and their biological reductant, flavodoxin.  相似文献   

14.
A peptide, N‐Ac‐Phe‐Tyr‐NH2, with angiotensin I‐converting enzyme (ACE) inhibitor activity was synthesized by an α‐chymotrypsin‐catalyzed condensation reaction of N‐acetyl phenylalanine ethyl ester (N‐Ac‐Phe‐OEt) and tyrosinamide (Tyr‐NH2). Three kinds of solvents: a Tris–HCl buffer (80 mM, pH 9.0), dimethylsulfoxide (DMSO), and acetonitrile were employed in this study. The optimum reaction solvent component was determined by simplex centroid mixture design. The synthesis efficiency was enhanced in an organic‐aqueous solvent (Tris‐HCl buffer: DMSO: acetonitrile = 2:1:1) in which 73.55% of the yield of N‐Ac‐Phe‐Tyr‐NH2 could be achieved. Furthermore, the effect of reaction parameters on the yield was evaluated by response surface methodology (RSM) using a central composite rotatable design (CCRD). Based on a ridge max analysis, the optimum condition for this peptide synthesis included a reaction time of 7.4 min, a reaction temperature of 28.1°C, an enzyme activity of 98.9 U, and a substrate molar ratio (Phe:Tyr) of 1:2.8. The predicted and the actual (experimental) yields were 87.6 and 85.5%, respectively. The experimental design and RSM performed well in the optimization of synthesis of N‐Ac‐Phe‐Tyr‐NH2, so it is expected to be an effective method for obtaining a good yield of enzymatic peptide. © 2012 American Institute of Chemical Engineers Biotechnol. Prog., 2012  相似文献   

15.
Recently, we have observed a nuclear localization for human α1‐antichymotrypsin (AACT) expressed in the cytosol of transgenic Bright Yellow‐2 (BY‐2) tobacco cultured cells (see accompanying paper: Benchabane, M., Saint‐Jore‐Dupas, C., Bardor, M., Faye, L., Michaud, D. and Gomord, V. (2008a) Targeting and post‐translational processing of human α1‐antichymotrypsin in BY‐2 tobacco cultured cells. Plant Biotechnol. J. doi: 10.1111/j.1467‐7652.2008.00382.x). In the present article, we assess whether the intrinsic DNA‐binding activity of AACT can explain its nuclear localization, and whether this same activity has an impact on its protease inhibitory potency and stability in planta. An engineered form of AACT with no DNA‐binding activity, rAACTΔK, was compared with the wild‐type polypeptide, rAACT, in terms of chymotrypsin inhibitory potency, stability in planta and distribution in tobacco cells. In accordance with available data reporting distinct sites for protease inhibition and DNA binding, rAACT and rAACTΔK showed similar antichymotrypsin activity, similar to the activity of native AACT purified from human plasma. As observed for AACT in BY‐2 tobacco cells, a green fluorescent protein (GFP)‐AACT fusion transiently expressed in the cytosol of tobacco leaf epidermal cells was detected mainly in the nucleus by confocal laser microscopy. By contrast, rAACTΔK expressed as a GFP fusion showed a balanced distribution between the cytosol and the nucleus, similar to the distribution pattern of free GFP exhibiting no DNA‐binding affinity. In line with immunodetection data showing higher accumulation levels for GFP‐AACT in tobacco leaf cells, rAACTΔK was more susceptible than rAACT to tryptic digestion in the presence of DNA. Overall, these observations suggest the following: (i) a retention effect of DNA on AACT in the nucleus; and (ii) a stabilizing effect of the AACT–DNA interaction on rAACT challenged with non‐target proteases, which, possibly, may be useful in protecting this protein in plant expression platforms.  相似文献   

16.
3,4‐Dihydroxy‐2‐butanone‐4‐phosphate synthase (DHBPS) encoded by ribB gene is one of the first enzymes in riboflavin biosynthesis pathway and catalyzes the conversion of ribulose‐5‐phosphate (Ru5P) to 3,4‐dihydroxy‐2‐butanone‐4‐phosphate and formate. DHBPS is an attractive target for developing anti‐bacterial drugs as this enzyme is essential for pathogens, but absent in humans. The recombinant DHBPS enzyme of Salmonella requires magnesium ion for its activity and catalyzes the formation of 3,4‐dihydroxy‐2‐butanone‐4‐phosphate from Ru5P at a rate of 199 nmol min?1 mg?1 with Km value of 116 μM at 37°C. Further, we have determined the crystal structures of Salmonella DHBPS in complex with sulfate, Ru5P and sulfate‐zinc ion at a resolution of 2.80, 2.52, and 1.86 Å, respectively. Analysis of these crystal structures reveals that the acidic loop (residues 34–39) responsible for the acid‐base catalysis is disordered in the absence of substrate or metal ion at the active site. Upon binding either substrate or sulfate and metal ions, the acidic loop becomes stabilized, adopts a closed conformation and interacts with the substrate. Our structure for the first time reveals that binding of substrate Ru5P alone is sufficient for the stabilization of the acidic active site loop into a closed conformation. In addition, the Glu38 residue from the acidic active site loop undergoes a conformational change upon Ru5P binding, which helps in positioning the second metal ion that stabilizes the Ru5P and the reaction intermediates. This is the first structural report of DHBPS in complex with either substrate or metal ion from any eubacteria. Proteins 2010. © 2010 Wiley‐Liss, Inc.  相似文献   

17.
The toxicity of mercury is often attributed to its tight binding to cysteine thiolate anions in vital enzymes. To test our hypothesis that Hg(II) binding to histidine could be a significant factor in mercury's toxic effects, we studied the enzyme chymotrypsin, which lacks free cysteine thiols; we found that chymotrypsin is not only inhibited, but also denatured by Hg(II). We followed the aggregation of denatured enzyme by the increase in visible absorbance due to light scattering. Hg(II)‐induced chymotrypsin precipitation increased dramatically above pH 6.5, and free imidazole inhibited this precipitation, implicating histidine‐Hg(II) binding in the process of chymotrypsin denaturation/aggregation. Diethylpyrocarbonate (DEPC) blocked chymotrypsin's two histidines (his40 and his57) quickly and completely, with an IC50 of 35 ± 6 µM. DEPC at 350 µM reduced the hydrolytic activity of chymotrypsin by 90%, suggesting that low concentrations of DEPC react with his57 at the active site catalytic triad; furthermore, DEPC below 400 µM enhanced the Hg(II)‐induced precipitation of chymotrypsin. We conclude that his57 reacts readily with DEPC, causing enzyme inhibition and enhancement of Hg(II)‐induced aggregation. Above 500 µM, DEPC inhibited Hg(II)‐induced precipitation, and [DEPC] >2.5 mM completely protected chymotrypsin against precipitation. This suggests that his40 reacts less readily with DEPC, and that chymotrypsin denaturation is caused by Hg(II) binding specifically to the his40 residue. Finally, we show that Hg(II)‐histidine binding may trigger hemoglobin aggregation as well. Because of results with these two enzymes, we suggest that metal‐histidine binding may be key to understanding all heavy metal‐induced protein aggregation.  相似文献   

18.
An α‐amylase and a glucoamylase produced by Thermomyces lanuginosus F1 were separated by ion‐exchange chromatography on Q‐Sepharose fast flow. The enzymes were further purified to electrophoretic homogeneity by chromatography on Sephadex G‐100 and Phenyl‐Sepharose CL‐4B.The molecular weights and isoelectric points of the enzymes were 55,000 Da and pHi 4.0 for α‐amylase and 70,000 Da and pHi 4.0 for glucoamylase, respectively. The optimum pH and temperatures for the enzymes were found to be 5.0 and 60 °C for α‐amylase, and 6.0 and 70 °C for glucoamylase,respectively. Both enzymes were maximally stable at pH 4.0 and retained over 80% of their activity between pH 5.0 and 6.0 for 24 h. After incubation at 90 °C (1 h), the α‐amylase and glucoamylase retained only 6% and 16% of their activity, respectively. The enzymes readily hydrolyzed soluble starch, amylose, amylopectin and glycogen but hydrolyzed pullulan very slowly. Glucoamylase and α‐amylase had highest affinity for soluble starch with KM values of 0.80 mg/ml and 0.67 mg/ml, respectively. The α‐amylase hydrolyzed raw starch granules with a predominant production of glucose and maltose. The activities of α‐amylase and glucoamylase increased in the presence of Mn2+, Co2+, Ca2+, Zn2+ and Fe2+, but were inhibited by guanidine‐HCl, urea and disodium EDTA. Both enzymes possess pH and thermal stability characteristics that may be of technological significance.  相似文献   

19.
Two selective agonists of nonopioid β‐endorphin receptor, synthetic peptides TPLVTLFK (octarphin) and SLTCLVKGFY (immunorphin), were labeled with tritium to specific activity of 29 and 25 Ci/mmol, respectively. Both labeled peptides were found to bind to high‐affinity naloxone‐insensitive binding sites on the membranes isolated from the rat myocardium (Kd = 2.0 ± 0.2 and 2.5 ± 0.3 nM, respectively). The [3H]octarphin specific binding to the myocardial membranes was inhibited by unlabeled β‐endorphin (Ki = 1.9 ± 0.2 nM) and immunorphin (Ki = 2.2 ± 0.3 nM). The [3H]immunorphin specific binding with the membranes was inhibited by unlabeled β‐endorphin (Ki = 2.3 ± 0.3 nM) and octarphin (Ki = 2.4 ± 0.3 nM). The binding specificity study revealed that these binding sites were insensitive not only to naloxone but also to α‐endorphin, γ‐endorphin, [Met5]enkephalin and [Leu5]enkephalin. Thus, β‐endorphin, immunorphin and octarphin bind to the common high‐affinity naloxone‐insensitive receptor of the rat myocardial membranes. Copyright © 2011 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

20.
5α‐Androst‐16‐en‐3α‐ol (α‐androstenol) is an important contributor to human axilla sweat odor. It is assumed that α‐andostenol is excreted from the apocrine glands via a H2O‐soluble conjugate, and this precursor was formally characterized in this study for the first time in human sweat. The possible H2O‐soluble precursors, sulfate and glucuronide derivatives, were synthesized as analytical standards, i.e., α‐androstenol, β‐androstenol sulfates, 5α‐androsta‐5,16‐dien‐3β‐ol (β‐androstadienol) sulfate, α‐androstenol β‐glucuronide, α‐androstenol α‐glucuronide, β‐androstadienol β‐glucuronide, and α‐androstenol β‐glucuronide furanose. The occurrence of α‐androstenol β‐glucuronide was established by ultra performance liquid chromatography (UPLC)/MS (heated electrospray ionization (HESI)) in negative‐ion mode in pooled human sweat, containing eccrine and apocrine secretions and collected from 25 female and 24 male underarms. Its concentration was of 79 ng/ml in female secretions and 241 ng/ml in male secretions. The release of α‐androstenol was observed after incubation of the sterile human sweat or α‐androstenol β‐glucuronide with a commercial glucuronidase enzyme, the urine‐isolated bacteria Streptococcus agalactiae, and the skin bacteria Staphylococcus warneri DSM 20316, Staphylococcus haemolyticus DSM 20263, and Propionibacterium acnes ATCC 6919, reported to have β‐glucuronidase activities. We demonstrated that if α‐ and β‐androstenols and androstadienol sulfates were present in human sweat, their concentrations would be too low to be considered as potential precursors of malodors; therefore, the H2O‐soluble precursor of α‐androstenol in apocrine secretion should be a β‐glucuronide.  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