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.
An α-glucosidase activity (EC 3.2.1.20) isolated from Sulfolobus solfataricus strain MT-4 was characterised and found of interest at industrial level in the saccharification step of hydrolysis process of starch. The gene encoding for the enzyme was expressed in Escherichia coli BL21 (DE3) with a yield of 87.5 U/g of wet biomass. The recombinant cytosolic enzyme was purified to homogeneity with a rapid purification procedure employing only steps of selective and progressive thermal precipitations with a final yield of 75.4% and a purification of 14.5-fold. The properties of this thermophilic α-glucosidase were compared with those of the α-glucosidase of a commercial preparation from Aspergillus niger used in the starch processing.  相似文献   

2.
In recent years several 15β-hydroxysteroids have emerged pathognomonic of adrenal disorders in human neonates of which 3α,15β,17α-trihydroxy-5β-pregnan-20-one (2) was the first to be identified in the urine of newborn infants affected with congenital adrenal hyperplasia. In this investigation we report the synthesis of the three remaining 3ξ,5ξ-isomers, namely 3α,15β,17α-trihydroxy-5α-pregnan-20-one (3), 3β,15β,17α-trihydroxy-5α-pregnan-20-one (7) and 3β,15β,17α-trihydroxy-5β-pregnan-20-one (8) for their definitive identification in pathological conditions in human neonates. 3β,15β-Diacetoxy-17α-hydroxy-5-pregnen-20-one (11), a product of chemical synthesis was converted to the isomeric 3 and 7, while conversion of 15β,17α-dihydroxy-4-pregnen-3,20-dione (4), a product of microbiological transformation, resulted in the preparation of 8. In brief, selective acetate hydrolysis of 11 gave 15β-acetoxy-3β,17α-dihydroxy-5-pregnen-20-one (12) which on catalytic hydrogenation gave 15β-acetoxy-3β,17α-dihydroxy-5α-pregnan-20-one (13) a common intermediate for the synthesis of the 3β(and α),5α-isomers. Hydrolysis of the 15β-acetate gave 7, whereas oxidation with pyridinium chlorochromate gave 15β-acetoxy-17α-hydroxy-5α-pregnan-3,20-dione (14) which on reduction with -Selectride and hydrolysis of the 15β-acetate gave 3. Finally, hydrogenation of 4 gave 15β,17α-dihydroxy-5β-pregnan-3,20-dione (10) which on reduction with -Selectride gave 8.  相似文献   

3.
An affinity chromatographic method with a novel eluant from Bacillus licheniformis is described. α-amylase was bound to starch, starch-celite, starch-Sepharose columns and the bound α-amylase was rapidly eluted with 2% (w/v) white dextrin. The binding capacity of α-amylase to starch column is 380 μmol/g of starch. The purified enzyme showed a single polypeptide on SDS-polyacrylamide gel electrophoresis with a molecular weight of 58 kD. The specificity of purified enzyme was confirmed by immunodiffusion, immunoelectrophoresis. Single radial immunodiffusion and western blotting studies analyzed the synthesis of enzyme at different time points.  相似文献   

4.
5.
Hydroquinone α-isomaltoside and hydroquinone α-glucoside were synthesized by transglucosylation in an aqueous system with baker's yeast α-glucosidase from hydroquinone and maltose as a glucosyl donor. Only one phenolic group was glucosylated, with α-selectivity, and the nature of the reaction products was governed by the concentration of hydroquinone. The optimal conditions for synthesis of glycosides were 9 mM hydroquinone and 1.5 M maltose in a 100 mM sodium citrate/phosphate buffer at pH 5.0 and 30 °C for 20 h. Under these conditions both hydroquinone α-glycosides were obtained in nearly equimolar amounts with a total molar yield of 28% with respect to hydroquinone and a total glycoside concentration of 1 mg/mL in the reaction mixture.  相似文献   

6.
β-Glucosidases (Glu1 and Glu2) in maize specifically interact with a lectin called β-glucosidase aggregating factor (BGAF). We have shown that the N-terminal (Glu50–Val145) and the C-terminal (Phe466–Ala512) regions of maize Glu1 are involved in binding to BGAF. Sequence comparison between sorghum β-glucosidases (dhurrinases, which do not bind to BGAF) and maize β-glucosidases, and the 3D-structure of Glu1 suggested that the BGAF-binding site on Glu1 is much smaller than predicted previously. To define more precisely the BGAF-binding site, we constructed additional chimeric β-glucosidases. The results showed that a region spanning 11 amino acids (Ile72–Thr82) on Glu1 is essential and sufficient for BGAF binding, whereas the extreme N-terminal region Ser1–Thr29, together with C-terminal region Phe466–Ala512, affects the size of Glu1–BGAF complexes. The dissociation constants (Kd) of chimeric β-glucosidase–BGAF interactions also demonstrated that the extreme N-terminal and C-terminal regions are important but not essential for binding. To confirm the importance of Ile72–Thr82 on Glu1 for BGAF binding, we constructed a chimeric sorghum β-glucosidase, Dhr2 (C-11, Dhr2 whose Val72–Glu82 region was replaced with the Ile72–Thr82 region of Glu1). C-11 binds to BGAF, indicating that the Ile72–Thr82 region is indeed a major interaction site on Glu1 involved in BGAF binding.  相似文献   

7.
Thermostable β-glucosidase from Sulfolobus shibatae was immobilized on silica gel modified or not modified with 3-aminopropyl-triethoxysilane using transglutaminase as a cross-linking factor. Obtained preparations had specific activity of 3883 U/g of the support, when measured at 70 °C using o-nitrophenyl β-d-galactopyranoside (GalβoNp) as substrate. The highest immobilization yield of the enzyme was achieved at pH 5.0 in reaction media. The most active preparations of immobilized β-glucosidase were obtained at a transglutaminase concentration of 40 mg/ml at 50 °C. The immobilization was almost completely terminated after 100 min of the reaction and prolonged time of this process did not cause considerable changes of the activity of the preparations. The immobilization did not influence considerably on optimum pH and temperature of GalβoNp hydrolysis catalyzed by the investigated enzyme (98 °C, pH 5.5). The broad substrate specifity and properties of the thermostable β-glucosidase from S. shibatae immobilized on silica-gel indicate its suitability for hydrolysis of lactose during whey processing.  相似文献   

8.
Folding type-specific secondary structure propensities of 20 naturally occurring amino acids have been derived from α-helical, β-sheet, α/β, and α+β proteins of known structures. These data show that each residue type of amino acids has intrinsic propensities in different regions of secondary structures for different folding types of proteins. Each of the folding types shows markedly different rank ordering, indicating folding type-specific effects on the secondary structure propensities of amino acids. Rigorous statistical tests have been made to validate the folding type-specific effects. It should be noted that α and β proteins have relatively small α-helices and β-strands forming propensities respectively compared with those of α+β and α/β proteins. This may suggest that, with more complex architectures than α and β proteins, α+β and α/β proteins require larger propensities to distinguish from interacting α-helices and β-strands. Our finding of folding type-specific secondary structure propensities suggests that sequence space accessible to each folding type may have differing features. Differing sequence space features might be constrained by topological requirement for each of the folding types. Almost all strong β-sheet forming residues are hydrophobic in character regardless of folding types, thus suggesting the hydrophobicities of side chains as a key determinant of β-sheet structures. In contrast, conformational entropy of side chains is a major determinant of the helical propensities of amino acids, although other interactions such as hydrophobicities and charged interactions cannot be neglected. These results will be helpful to protein design, class-based secondary structure prediction, and protein folding. © 1998 John Wiley & Sons, Inc. Biopoly 45: 35–49, 1998  相似文献   

9.
Binding of receptor-recognized forms of tetrameric human α2-macroglobulin (α2M*) to a macrophage signaling receptor induces cAMP synthesis, increases in inositol 1,4,5-triphosphate (IP3) synthesis, and a concomitant rise in cytosolic free calcium ([Ca2+]i). The α2M* signaling receptor is coupled to a pertussis-toxin insensitive G protein. Binding of α2M* also occurs to the low density lipoprotein receptor-related protein/α2M receptor (LRP/α2MR), but this binding does not induce signal transduction. Rat α1-inhibitor-3 (α1I3) is a monomeric member of the α-macroglobulin/complement superfamily. Like α2M, it can react with proteinases or methylamine which induces a conformational change causing activated α1I3 to bind to LRP/α2MR. We now report that α1I3-methylamine binds to the macrophage α2M* signaling receptor inducing a rapid rise in the synthesis of IP3 with a subsequent 1.5- to 3-fold rise in [Ca2+]i. α1I3-methylamine binding to macrophages also caused a statistically significant elevation in cAMP. Native α1I3, like α2M, was unable to induce signal transduction. α1I3 forms a complex with α1-microglobulin, which has a distinct conformation from α1I3 and is recognized by LRP/α2MR. This complex also induces an increase in [Ca2+]i comparable to the effect of α1I3-methylamine on macrophages. It is concluded that activation of α1I3 by methylamine or binding of α1-microglobulin causes similar conformational changes in the inhibitor, exposing the receptor recognition site for the α2M* signaling receptor, as well as for LRP/α2MR. © 1996 Wiley-Liss, Inc.  相似文献   

10.
In this paper, bis (indol-3-yl) methanes (BIMs) were synthesised and evaluated for their inhibitory activity against α-glucosidase and α-amylase. All synthesised compounds showed potential α-glucosidase and α-amylase inhibitory activities. Compounds 5 g (IC50: 7.54 ± 1.10 μM), 5e (IC50: 9.00 ± 0.97 μM), and 5 h (IC50: 9.57 ± 0.62 μM) presented strongest inhibitory activities against α-glucosidase, that were ∼ 30 times stronger than acarbose. Compounds 5 g (IC50: 32.18 ± 1.66 µM), 5 h (IC50: 31.47 ± 1.42 µM), and 5 s (IC50: 30.91 ± 0.86 µM) showed strongest inhibitory activities towards α-amylase, ∼ 2.5 times stronger than acarbose. The mechanisms and docking simulation of the compounds were also studied. Compounds 5 g and 5 h exhibited bifunctional inhibitory activity against these two enzymes. Furthermore, compounds showed no toxicity against 3T3-L1 cells and HepG2 cells.

Highlights

  1. A series of bis (indol-3-yl) methanes (BIMs) were synthesised and evaluated inhibitory activities against α-glucosidase and α-amylase.
  2. Compound 5g exhibited promising activity (IC50 = 7.54 ± 1.10 μM) against α-glucosidase.
  3. Compound 5s exhibited promising activity (IC50 = 30.91 ± 0.86 μM) against α-amylase.
  4. In silico studies were performed to confirm the binding interactions of synthetic compounds with the enzyme active site.
  相似文献   

11.
α-Ketobutyrate decarboxylase encoded in the -methionine catabolism operon of Pseudomonas putida is homologous with the E1 component of pyruvate dehydrogenase complex from gram-negative bacteria. The enzyme was purified to homogeneity from the cell extract of an Escherichia coli transformant. The purified enzyme was homodimeric with a subunit of Mr 93,000 on SDS-PAGE. The enzyme activity was activated by the addition of both thiamine pyrophosphate (TPP) and a divalent cation, such as Mg2+, Mn2+, and Co2+. The enzyme showed high activity for α-ketobutyrate and α-keto-n-valerate rather than pyruvate, but the α-keto acids with increasing length of the side chain as well as branching, such as α-keto-n-caproate and α-keto-3-methylvalerate, were not used by the enzyme. The Km values for α-ketobutyrate and pyruvate were 0.016 and 0.147 mM, respectively, and the kcat/Km value (10.69 s−1 mM−1) for α-ketobutyrate was 29-fold greater than that for pyruvate. Thus, α-ketobutyrate decarboxylase is distinguished from the pyruvate dehydrogenase E1 component with respect to the substrate specificity, although their structural and enzymological properties were similar. These results suggest that the unique substrate specificity of α-ketobutyrate decarboxylase is due to a slight difference in the highly conserved active sites of both enzymes.  相似文献   

12.
The production of extracellular α-amylase by thermotolerant Bacillus subtilis was studied in solid state fermentation (SSF). The effect of wheat bran (WB) and rice husk (RH) was examined. The appropriate incubation period, moisture level, particle size and inoculum concentration was determined. Maximum yields of 159,520 and 21,760 U g−1 were achieved by employing WB and RH as substrates in 0.1 M phosphate buffer at pH 7 with 30% initial moisture content at 24 and 48 h. Particle size and inoculum concentration were found to be 1000 μm, 20% and 500 μm, 15% for WB and RH, respectively. Enzyme yield was 7.3-fold higher with WB medium compared with RH.  相似文献   

13.
Prostaglandin H synthase (PHS) has gained interest as a drugmetabolizing enzyme and has been shown to cooxidize and metabolically activate diethylstilbestrol (DES) in vitro. Both 7,8-benzoflavone (α-naphthoflavone, ANF) and 5,6-benzoflavone (β-naphthoflavone, BNF) have now been studied for their effects on PHS from ram seminal vesicle microsomes by means of several in vitro assays. The PHS-catalyzed cooxidation of DES, as measured by high-performance liquid chromatography (HPLC) analysis, is inhibited by BNF and ANF at micromolar concentrations, with median inhibitory concentrations (IC-50) of<20 and 40 μM, respectively. The oxidation of DES is inhibited whether it is initiated by arachidonic acid or by hydrogen peroxide, indicating that the benzoflavones inhibit PHS by a mechanism different from that of indomethacin. Monitoring of cyclooxygenase activity in an oxygraph also reveals an inhibition of PHS by BNF which depends only weakly on arachidonic acid concentration; inhibition by ANF is less pronounced under these conditions. Since PHS-catalyzed conversion of the benzoflavone compounds was detected under conditions permitting cooxidation, the inhibition of PHS by benzoflavones in vitro could either be a direct effect or possibly mediated via metabolites. Our data imply that ANF and BNF, in addition to their well-known role as modifiers of mixed-function oxidases, can affect the PHS-catalyzed metabolism of xenobiotics. This is discussed in the context of adverse effects caused by DES in vivo and in cell culture and must be taken into account when interpreting the modifying effect of benzoflavones on these endpoints.  相似文献   

14.
The glgP gene encoding α-glucan phosphorylase (α-GP) from the thermopile Thermus caldophilus GK24 has been identified, cloned, and overexpressed in Escherichia coli and used to synthesize d-glucose-1-phospate (G1P) from an inexpensive starch. The enzyme, purified 6.5-fold, was isolated in 31% yield from the transformed E. coli, and gave a single band. The purified enzyme may exist as a homohexamer with an apparent molecular mass of a 550 kDa molecule, consisting of 90 kDa per subunit. The optimal pH and temperature were 7.0 and 70 °C in the α-GP reaction with starch producing G1P. Soluble starch (amylopectin, amylose) turned out to be a better substrate giving a higher yield of G1P than α-1,6-branched α-1,4-glucans (glycogen, potato starch, etc.). As a result, G1P was obtained in a good yield (47%, w/w) from the reaction containing 5% (w/v) soluble starch in 0.7 M potassium phosphate at pH 7.0. T. caldophilus α-GP shows a high tolerance (up to 0.7 M) of potassium phosphate and plays a critical role in shifting the reaction equilibrium in favor of G1P synthesis. The G1P product can be purified simply by ethanol precipitation, after removing the unreacted starch and inorganic phosphate by activated charcoal and magnesium acetate precipitation. It is concluded that T. caldophilus α-GP readily utilized in large scale synthesis of G1P.  相似文献   

15.
The maturation of connective tissue involves the organization of collagen fibres by resident fibroblasts. Fibroblast attachment to collagen has been demonstrated to involve cell surface receptors, integrins of the β1 family. Integrins are associated with cytoplasmic actin of microfilaments either directly or through focal adhesions. The major actin isoform of fibroblast microfilaments is β actin and to a lesser extent α smooth muscle (α SM) actin. Cultured human dermal fibroblasts derived from adult dermis, newborn foreskin or keloid scar were grown on either uncoated or collagen-coated surfaces. The expression and synthesis of both α2β1 integrin and α SM actin were followed by immunohistology and immunoprecipitation. Fibroblasts on uncoated surfaces expressed little α2β1 integrin on their surface, while 20 per cent of them demonstrated α SM actin within microfilaments. Fibroblasts grown on a collagen-coated surface minimally expressed α SM actin in microfilament structures and a majority of the cells were positive for α2β1 integrin on their membranes. Using [35S]-methionine incorporation and immunoprecipitation, it was shown that fibroblasts grown in uncoated dishes synthesized more α SM actin than fibroblasts grown on collagen-coated dishes. In contrast, fibroblasts grown on collagen coated dishes synthesized more α2β1 integrin compared to the same cells grown on uncoated dishes. Fibroblasts maintained on a type I collagen upregulate the expression and synthesis of α2β1 integrin, and downregulate the expression and synthesis of α SM actin. © 1998 John Wiley & Sons, Ltd.  相似文献   

16.
For efficient alkyl glucoside production from cellooligosaccharides, we constructed a yeast strain for alkyl glucoside synthesis by genetically inducing the display of β-glucosidase 1 (BGL1) from the filamentous fungus Aspergillus aculeatus No. F-50 on the cell surface. The localization of BGL1 on the cell surface was confirmed by immunofluorescence microscopy. The yeast strain displaying BGL1 catalyzed alkyl glucoside synthesis from p-nitrophenyl β-d-glucoside and primary alcohols. The highest yield of alkyl glucoside was 27.3% of the total sugar. The substrate specificities of the BGL1-displaying yeast strain and almond β-glucosidase were compared using different-chain-length cellooligosaccharides. The BGL1-displaying yeast showed efficient alkyl glucoside production from not only glucose but also cellohexaose. This yeast is applicable as a whole-cell biocatalyst for alkyl glucoside production from cellulose hydrolysates.  相似文献   

17.
The metabolism of the illegal growth promoter ethylestrenol (EES) was evaluated in bovine liver cells and subcellular fractions of bovine liver preparations. Incubations with bovine microsomal preparations revealed that EES is extensively biotransformed into norethandrolone (NE), another illegal growth promoter. Furthermore, incubations of monolayer cultures of hepatocytes with NE indicated that NE itself is rapidly reduced to 17α-ethyl-5β-estrane-3α,17β-diol (EED). In vivo tests confirmed that, after administration of either EES or NE, EED is excreted as a major metabolite. Therefore, it was concluded that, both in urine and faeces samples, EED can be used as a biological marker for the illegal use of EES and/or NE. Moreover, by monitoring EED in urine or faeces samples, the detection period after NE administration is significantly prolonged. These findings were further confirmed by three cases of norethandrolone abuse in a routine screening program for forbidden growth promoters.  相似文献   

18.
Cardiomyocyte tumour necrosis factor α (TNF‐α) production contributes to myocardial depression during sepsis. This study was designed to observe the effect of norepinephrine (NE) on lipopolysaccharide (LPS)‐induced cardiomyocyte TNF‐α expression and to further investigate the underlying mechanisms in neonatal rat cardiomyocytes and endotoxaemic mice. In cultured neonatal rat cardiomyocytes, NE inhibited LPS‐induced TNF‐α production in a dose‐dependent manner. α1‐ adrenoceptor (AR) antagonist (prazosin), but neither β1‐ nor β2‐AR antagonist, abrogated the inhibitory effect of NE on LPS‐stimulated TNF‐α production. Furthermore, phenylephrine (PE), an α1‐AR agonist, also suppressed LPS‐induced TNF‐α production. NE inhibited p38 phosphorylation and NF‐κB activation, but enhanced extracellular signal‐regulated kinase 1/2 (ERK1/2) phosphorylation and c‐Fos expression in LPS‐treated cardiomyocytes, all of which were reversed by prazosin pre‐treatment. To determine whether ERK1/2 regulates c‐Fos expression, p38 phosphorylation, NF‐κB activation and TNF‐α production, cardiomyocytes were also treated with U0126, a selective ERK1/2 inhibitor. Treatment with U0126 reversed the effects of NE on c‐Fos expression, p38 mitogen‐activated protein kinase (MAPK) phosphorylation and TNF‐α production, but not NF‐κB activation in LPS‐challenged cardiomyocytes. In addition, pre‐treatment with SB202190, a p38 MAPK inhibitor, partly inhibited LPS‐induced TNF‐α production in cardiomyocytes. In endotoxaemic mice, PE promoted myocardial ERK1/2 phosphorylation and c‐Fos expression, inhibited p38 phosphorylation and IκBα degradation, reduced myocardial TNF‐α production and prevented LPS‐provoked cardiac dysfunction. Altogether, these findings indicate that activation of α1‐AR by NE suppresses LPS‐induced cardiomyocyte TNF‐α expression and improves cardiac dysfunction during endotoxaemia via promoting myocardial ERK phosphorylation and suppressing NF‐κB activation.  相似文献   

19.
Simulated enzymic debranching of a β-limit dextrin model, prepared from a computed construct made by random extension and branching, and given the CCL value of w-maize amylopectin (and equal amounts of external chains with ECL values of 2 and 3) has been related to experimental chromatograms of the debranched β-limit dextrin of the amylopectin. The profile was similar to those from gel chromatograms and IEC-PAD chromatography.The equivalent lengths in glucosyl units of grid-links (g-links) of internal and external chains in constructs were calculated from the ICL and ECL values of amylopectin and models produced from the constructs with the appropriate lengths for internal and external chains. These derived models were subjected to simulated hydrolysis by Pseudomonas stutzeri amylase and the products compared with those of the experimental distribution from w-maize amylopectin. With the model the amounts of maltotetraose and maltodextrins released were similar to the experimental values but the distribution of branched maltodextrins was quite different. Unlike w-maize amylopectin – a polymer with the cluster structure – which has given a profile of molecular sizes of maltodextrins with low amounts of single and small numbers of internal chains and with a peak at a MW of about 14,000 (13 chains), in the model the proportion of maltodextrin with one internal chain was high and as d.p. increased the amounts decreased exponentially. This would be expected if the distribution of internal chains in the core was random. It is suggested that in the core of a model prepared from a construct made with alternating probabilities of extension – one in which this probability is high relative to branching, and a second in which it is low – may give clusters of branched maltodextrins with short internal chains which are joined by longer chains; more closely approximating the distribution of internal chains of different lengths in amylopectin.An arrangement for amylopectin molecules in the starch granule has been proposed. In this, they have a wafer-like, discoidal shape, composed of the amorphous zone overlain with the double helical, crystalline region. The flat macromolecules are concentrically layered with the former on the inside and the latter oriented to the outside of the granule.  相似文献   

20.
Inflammatory cytokines are closely related to pigmentary changes. In this study, the effects of IFN‐γ on melanogenesis were investigated. IFN‐γ inhibits basal and α‐MSH‐induced melanogenesis in B16 melanoma cells and normal human melanocytes. MITF mRNA and protein expressions were significantly inhibited in response to IFN‐γ. IFN‐γ inhibited CREB binding to the MITF promoter but did not affect CREB phosphorylation. Instead, IFN‐γ inhibited the association of CBP and CREB through the increased association between CREB binding protein (CBP) and STAT1. These findings suggest that IFN‐γ inhibits both basal and α‐MSH‐induced melanogenesis by inhibiting MITF expression. The inhibitory action of IFN‐γ in α‐MSH‐induced melanogenesis is likely to be associated with the sequestration of CBP via the association between CBP and STAT1. These data suggest that IFN‐γ plays a role in controlling inflammation‐ or UV‐induced pigmentary changes.  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