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.
Human α1-antitrypsin (α1-AT) is a major serine protease inhibitor in plasma, secreted as a glycoprotein with a complex type of carbohydrate at three asparagine residues. To study glycosylation of heterologous proteins in yeast, we investigated the glycosylation pattern of the human α1-AT secreted in the baker's yeast Saccharomyces cerevisiae and in the methylotrophic yeasts, Hansenula polymorpha and Pichia pastoris. The partial digestion of the recombinant α1-AT with endoglycosidase H and the expression in the mnn9 deletion mutant of S. cerevisiae showed that the recombinant α1-AT secreted in S. cerevisiae was heterogeneous, consisting of molecules containing core carbohydrates on either two or all three asparagine residues. Besides the core carbohydrates, variable numbers of mannose outer chains were also added to some of the secreted α1-AT. The human α1-AT secreted in both methylotrophic yeasts was also heterogeneous and hypermannosylated as observed in S. cerevisiae, although the overall length of mannose outer chains of α1-AT in the methylotrophic yeasts appeared to be relatively shorter than those of α1-AT in S. cerevisiae. The α1-AT secreted from both methylotrophic yeasts retained its biological activity as an elastase inhibitor comparable to that of α1-AT from S. cerevisiae, suggesting that the different glycosylation profile does not affect the in vitro activity of the protein. © 1998 John Wiley & Sons, Ltd.  相似文献   

2.
The investigation focuses on the computer-aided generation of the molecular geometries, contact surfaces, and lipophilicity patterns of per-O-methylated α-CD ( 1 ) and its β-CD homolog 2 , and compares them with their parent non-substituted cyclodextrins. The molecular geometries, compared via statistical analysis of crystal structure data available, reveal 1 and 2 to be considerably more flexible than α-and β-CD, allowing wide variations in the tilting of the glucose units relative to the macrocyclic ring axes. The comparative evaluation of their contact surfaces not only discloses a substantial increase of the torus heights upon per-O-methylation (from ∼8.0Å in α- and β-CD, to → 11.1Å in 1 and 2 ), but also an enlargement of their cavity areas by 40% (+35Å2 for α-CD → 1 ) and 70% (+75Å2 for β-CD → 2 ), respectively. The hydrophobic characteristics of 1 and 2 , emerging from the molecular lipophilicity patterns (MLPs) generated and projected onto the contact surfaces in color-coded from, are inverse to those for α- and β-CD: the most hydrophobic surface regions of 1 and 2 are located at the torus rims made up by the 2-OMe and 3-OMe groups at one side, and the 6-CH2OMe moieties at the other, with a hydrophobic „band”︁ wrapping around the outside of the macrocycles; these „exo-lipophilic”︁ topographies are opposed by pronouncedly hydrophilic central cavities. A variety of experimental findings can be rationalized on the basis of the opposite lipophilicity profiles of the CDs and their permethylated analogs, such as for example the opposite orientation of benzaldehyde, p-nitrophenol, and 3-iodopropionic acid in the cavities of α-CD and of 1 . Thus, the notion is substantiated that the operation of dispersive interactions between guest and CD-host cavities play a more dominant role in inclusion complex formation than hitherto appreciated.  相似文献   

3.
Extracellular α-amylases I and II, produced by a facultative thermophile Bacillus thermoamyloliquefaciens KP 1071 capable of growing at 30–66°C, were purified to homogeneity. α-Amylase I consisted of a single polypeptide with methionine residue at the NH2-terminus. α-Amylase II consisted of two equivalent polypeptides each comprising a methionine at the NH2-terminus. α-Amylase I hydrolyzed endotypically α-1,4-bonds in glycogen, amylopectin and β-limit dextrin, but not their α-1,6-bonds. α-Amylase II degraded amylopectin and β-limit dextrin in exo-fashion by cleaving preferentially α-maltose units from the non-reducing ends and hydrolyzing their α-1,6-branch points. α-Amylase II hydrolyzed maltotriose, phenyl-α-maltoside, α- and β-cyclodextrins and pullulan, whereas α-amylase I had no activity for all these sugars. α-Amylases I and II hydrolyzed maltotetraose, maltopentaose, α-limit dextrin and amylose, but they were inactive for maltose, isomaltose and panose. It was suggested that α-amylase I is the most thermostable type of hitherto known maltotriogenic endo-acting α-amylases, and α-amylase II is the first maltogenic exo-acting α-amylase able to split α-1,6-bonds in amylopectin.  相似文献   

4.
Protein α-amylase inhibitors were prepared from wheat and their effects tested against insect storage pests both in vitro against the insect α-amylases and in vivo in insect feeding trials. Inhibitor fraction A was found to inhibit porcine pancreatic α-amylase but not insect α-amylases, whereas fractions B, C and D (0.28) did not inhibit porcine pancreatic α-amylase but were strong inhibitors of digestive α-amylases from larvae of Tribolium confusum, a storage pest of wheat products, and Callosobruchus maculatus, a storage pest of legume seeds. Fraction D, which was a single polypeptide of Mr 13 000 was the most effective inhibitor in vitro. It would appear that the degree of inhibition by the wheat α-amylase inhibitor preparations can be correlated with the presence of the Mr 13 000 (0.28) polypeptide since the purer this polypeptide the stronger was the inhibition; fraction A which contained two polypeptides of Mr 60 000 and 58 000 caused no inhibition. The effects of fractions B and C on larval development were determined in insect feeding trials. With C. maculatus both fractions were toxic, their relative effectiveness being directly paralleled by their effectiveness observed in vitro. Only fraction C was tested against T. confusum in feeding trials. Despite this fraction being equally effective against both pests in vitro it had very little effect upon larval development of T. confusum in vivo, thus suggesting that this organism is able to detoxify the wheat α-amylase inhibitors. As far as the authors are aware, this is the first time that the effects of identified inhibitor fractions have been monitored both in vitro and in vivo. The results, in contrast to previous proposals, suggest that selecting wheat varieties for high α-amylase inhibitory activity may not be a very reliable criterion in selecting for insect resistance.  相似文献   

5.
α-Glucosidase and α-glucoside permease are synthesized simultaneously in two different strains of baker's yeast (Saccharomyces cerevisiae) when the cells are induced with maltose. α-Thioethyl D-glucopyranoside (α-TEG) inhibits the transport of maltose into the cells, and the transport of maltose and α-TEG appeared to be mediated by the same permease system. There is an obvious correlation between α-glucosidase, α-glucoside permease and maltose fermentation activities in the yeasts while no correlation between these and the leavening ability of the yeasts can be demonstrated. Apparently the glucose concentration in dough is high enough to inhibit the permease-mediated transport of maltose into the cells thus impairing leavening ability from the maltose fermenting system.  相似文献   

6.
The action of α-amylases on β-cyclodextrin and the evidence of foreign activity of α-amylase in selected preparations of enzymes The interaction between cyclodextrins and α-amylases taken from different sources is discribed contradictious in the literature. Some α-amylases e.g. isolated from Aspergillus oryzae, porcine pancreas and saliva hydrolized cyclodextrins to glucose. The hydrolysis of cyclodextrins catalysed by α-amylase from Bacillus species have been described conflicting. In this paper the action of hydrolysis of different preparations of α-amylases on β-cyclodextrin have been investigated. It has been shown that Rohalase M3 (α-amylase from Aspergillus niger) cleaves the ring of β-cyclodextrin. 2 α-amylases from Bacillus subtilis are not able to hydrolyse β-cyclodextrin. The reasons for the different actions of hydrolysis have been discussed with size and structure of the active centre of the enzymes. Moreover, different preparation of hydrolysis have been tested on secondary activity of α-amylase. 2 glucoamylases from Aspergillus oryzae have been shown secondary activity of α-amylase. With the hydrolases α-glucosidase from fungies, β-amylase from malt, saccharase from yeast, invertase from S. cerevisiae and pullulanase from Aerobacter aerogenes no cleavage of the ring of β-cyclodextrin could be detected.  相似文献   

7.
Leaf of Ampelopsis grossedentata is a new resource of functional foods with healthful properties. Antioxidant and αglucosidase inhibitory activities of water extract (made in the style of drinking), tannin fraction (TF) and dihydromyricetin (DMY) from A. grossedentata leaves were evaluated. The main component of TF was identified as gallotannins. DPPH and ABTS radical scavenging activities and reducing power of TF were superior to those of water extract, however, inferior to those of DMY. In no PBS wash protocol of cellular antioxidant activity assay, DMY and TF exhibited similarly, while in PBS wash protocol, the value of TF was higher than that of DMY. In addition, TF possessed the highest αglucosidase inhibitory activities (IC50 = 1.94 μg mL?1), followed by water extract (IC50 = 23.10 μg mL?1) and DMY (IC50 = 72.21 μg mL?1). The strong αglucosidase inhibitory activity of TF may attribute to the binding capacity to enzymes, as confirmed by fluorescence analysis.  相似文献   

8.
L. Tarhan 《Starch - St?rke》1989,41(8):315-318
α-Amylase of Bacillus subtilis was immobilized on acrolein-styren copolymers of spongy and non-spongy structures which had been modified with hexamethylenediamine followed by glutaraldehyde. Optimum bonding rates were determined from the immobilization experiments carried out with different amounts of α-amylase, pH-activity relationships, variation of reaction rates with the time due to the substrate diffusion limitation factor and the observed Km values were evaluated.  相似文献   

9.
Sorghum malt α‐amylase can compete with bacterial α‐amylase in industrial applications, if sufficiently stable and produced in a large enough quantity. Conditions for maximal α‐amylase production in sorghum malt and the physico‐chemical properties of the α‐amylase so produced are reported in this study. Sorghum grains were steeped in buffers with varying pH (4.0–8.0) for 24 h, at room temperature, and germinated for another 48 h to obtain the green malt. The buffer that induced the highest quantity of α‐amylase was chosen as the optimal pH and served as the medium for further steeping experiments conducted at different temperatures (10, 20, 30, 40, 50 and 60°C). The α‐amylase activity in the extract was determined in order to obtain the optimum temperature for α‐amylase induction at this particular pH. For the purpose of comparison, the α‐amylase produced at the optimum pH and temperature was purified to apparent homogeneity by a combination of ion‐exchange and size‐exclusion chromatography, and further characterized. Eight‐fold higher α‐amylase activity was induced in pH 6.5 buffer at 20°C compared with water, the traditional steeping medium. The Km and Vmax were estimated to be 1.092 ± 0.05 mg mL?1 and 3516 ± 1.981 units min?1, respectively. The activation energy of the purified amylase for starch hydrolysis was 6.2 kcal K?1 mol?1. Chlorides of calcium and manganese served as good activators, whereas CuSO4 inhibited the enzyme with a 42% loss in activity at 312 mm salt concentration. Copyright © 2012 The Institute of Brewing & Distilling  相似文献   

10.
Treatment of glucosyl-α-cyclodextrin with radioactive glucose in the presence of Bacillus macerans cyclodextrin glucanotransferase, yielded radioactive maltose, maltotriose, and four kinds of glucosyl branched oligosaccharides. The branched oligosaccharides formed had identical Rf values as those of branched oligosaccharides of 5-8 glucose unit on paper chromatography using the solvent system of 1-butanol: pyridine: water (6/4/4), (Fractions B5-B8). Individual oligosaccharides were isolated and their structures determined using the action of porcine pancreatic α-amylase, β-amylase, pullulanase, isopullulanase and glucoamylase. The structure of the main component of the fraction B5 was 64-O-α-glucosylmaltotetraose, B6 was 64-O-α-glucosylmatopentaose, B7 was 65-O-α-glucosylmaltohexaose, and B8 was 66-O-α-glucosylmaltoheptaose. Fraction B8, which formed at the initial stage of the reaction, contained 64, 65 and 66-O-α-glucosylmaltoheptaose. Reducing end 14C-labelled 64-O-α-glucosylmaltotetraose was degraded to 63-O-α-glucosylmaltotriose and radioactive glucose by the exhaustive action of Bacillus macerans cycloamylose glucanotransferase, and 63-O-α-glucosylmaltotriose was completely non-reactive to the enzyme. From these results, we proposed a model of the active site of the enzyme.  相似文献   

11.
Some kinetic characteristics and hydrolytic action patterns on various β-D -galactosyl-maltooligosaccharides (Gal-Gn), ranging in size from D.P. (degree of polymerization) 5 to 8, of an exo-maltotetraose-forming amylase of Pseudomonas stutzeri (G4-amylase) were examined to produce a few p-nitrophenyl β-D -galactosyl-α-maltooligosaccharides (Gal-GnP, n = 4,5). The relative hydrolytic reaction rates for larger Gal-Gn by the enzyme were larger than those for smaller saccharides tough the values for unmodified linear maltooligosachharides were almost same. Michaelis constants (Km) for hydrolysis of Gal-G4, Gal-G5, Gal-G6 and Gal-G7 by the enzyme were 1.3, 1.9, 1.3 and 1.3mM, and apparent molecular activities (ko) for these saccharides were 5.9, 38, 91 and 126s−1, respectively. The values of ko/Km for them were remarkably smaller than those for unmodified linear maltooligosaccharides. The G4-amylase cleaved 2 points of the α-1,4-glucosidic linkage in β-1,4-Gal-G4 to give β-1,4-Gal-G2 and -Gal-G3 in the molar ratio of 3:1, whereas the enzyme attacked 3 points of the linkage in β-1,4-Gal-G5, -Gal-G6 and -Gal-G7 to form β-1,4-Gal-G2,-Gal-G3 and Gal-G4 in the molar ratios of 2:5:1, 1:3:6 and 1:3:6, in the early stage of the reaction, respectively. On the other hand, the enzyme showed no action on β-1,6-Gal-G4 and formed β-1,6-Gal-G4 solely from β-1,6-Gal-G5, and β-1,6-Gal-G4 and -Gal-G5 were from β-1,6-Gal-G6 and -Gal-G7 in the ratios of 8:1 and 2:1, respectively. The enzyme also catalyzed the transfer action to produce Gal-G3P, Gal-G4P and Gal-G5P, of which the formation ratio was coincided well with the hydrolytic action pattern on each Gal-Gn, from Gal-Gn tested as a donor and p-nitrophenyl α-glucoside (GP) as an acceptor in an aqueous solution containing 40% (v/v) methanol. By using this novel reaction, Gal-G5P is now producing on an industrial scale to apply as a substrate for the assay of human α-amylase.  相似文献   

12.
Changes of viscosity characteristics of genetically diverse hexaploid wheat starches during pasting in water, 1% NaCl solution, or at pH 4 or pH 10 were studied using a Rapid Visco‐Analyzer. Peak viscosity (PV( hot paste viscosity (HPV) and cool paste viscosity (CPV) of all the wheat starches was little affected in pH 4 and pH 10 treatments. In 1% NaCl, all starches showed substantial increases in all three parameters relative to pasting in water. The use of 1% β‐cyclodextrin (β‐CD) (cycloheptaamylose) solution increased PV of high‐swelling starches, but generally slightly decreased that of low swelling starches in all treatment conditions. HPV was always reduced by addition of 1% β‐CD, but CPV was increased in most treatments for Anza and Yecora Rojo (low swelling( but decreased for Klasic (high swelling). Bacterial α‐amylase was added to starch or flour. The effect of β‐CD was shown to be independent of α‐amylase inhibition in wheat starch, but β‐CD strongly inhibited α‐amylase in wheat flour.  相似文献   

13.
Here we present a comparative study of caprine β- and αs1-caseins behaviours at the air–water interface and in solution. Both caseins were purified from the milk of a single goat homozygous at the αs1- and β-Cn loci, with a high degree of purity (98%). Physical measurements (ellipsometry, surface pressure and surface rheology) were performed at the air–water interface, whereas SAXS measurements were performed on casein solutions. Our results clearly show that self-organizations, both at the air–water interface and in solution are different for β- and αs1-caseins. β-casein is unfolded in solution and forms a network at the interface, while αs1-casein forms compact objects in solution and is organised in fluid domains at the interface. We also show that the presence of Ca2+ in the subphase strongly disturbs the interfacial layer formed by the caseins. It is elsewhere worth noting that in solution, the aggregation of αs1-casein induced by calcium ions is associated with a pronounced change in the molecular structural organisation of the protein, which seems to adopt, in these conditions, an unfolded structure.  相似文献   

14.
Ten compounds were isolated and purified from the peels of gold‐red apple (Malus domestica) for the 1st time. The identified compounds are 3β, 20β‐dihydroxyursan‐28‐oic acid (1), 2α‐hydroxyoleanolic acid (2), euscaphic acid (3), 3‐O‐p‐coumaroyl tormentic acid (4), ursolic acid (5), 2α‐hydroxyursolic acid (6), oleanolic acid (7), betulinic acid (8), linolic acid (9), and α‐linolenic acid (10). Their structures were determined by interpreting their nuclear magnetic resonance and mass spectrometry (MS) spectra, and by comparison with literature data. Compound 1 is new, and compound 2 is herein reported for the 1st time for the genus Malus. α‐Glucosidase inhibition assay revealed 6 of the triterpenoid isolates as remarkable α‐glucosidase inhibitors, with betulinic acid showing the strongest inhibition (IC50 = 15.19 μM). Ultra‐performance liquid chromatography‐electrospray ionization MS analysis of the fruit peels, pomace, flesh, and juice revealed that the peels and pomace contained high levels of triterpenes, suggesting that wastes from the fruit juice industry could serve as rich sources of bioactive triterpenes.  相似文献   

15.
The temperature-dependent dyeability of oligoglucanes and polyglucanes with I2/KI solution is based on the interaction between iodine and the C-atoms of glucosidic bonds (Vetter and Thorn [1, 2]). Only when the helix is compressed, the expansion of the lumen that is necessary for the attachment of iodine bands is achieved as suggested by Freudenberg et al. in 1939 [3]. The development of colour starts with the maltododecaosis (just over two spirals) within a temperature range of 20 to 25°C adsorbing one pentaiodide and bringing about a pale shade of pink. The rich blue tone is achieved with pure amyloses forming a maximum of about 610 to 620 nm. Unramified polysaccharides which had been synthesised by our team from 30 glucoside residues, still showed a purple colour with a maximum of 510 nm at temperatures from 20 to 25°C. Defined polysaccharides from 40 residues are bluish-purple (just 7 spirals of a tension-free helix) with a maximum absorption of about 550 nm. According to our knowledge, the rich blue tone requires at least 50 glucoside residues linked in an unramified chain [1, 2]. Amylopectin from potatoes has a maximum of 575 to 580 nm which shifts to 555 to 560 nm after partial break-down into β-dextrin by β-amylase (EC 3.2.1.2) and does a long-wave re-shift to 570 nm after splitting off of the side-chain stumps in the basic structure by pullulanase (EC 3.2.1.41). In the case of maize, the maximum is 535 nm for amylopectin, 520 nm for β-dextrin, and 555 nm for the basic structure. This means that in amylopectin from potatoes the helical sections available for the adsorption of iodine bands are longer than those in amylopectin from maize [4]. In glycogen iodine accumulates in a diffuse manner without forming any long bands. For this reason, an absorption shoulder of merely 400 to 500 nm is found by photometry [4]. A treatment with β-amylase makes the situation even worse; only when the side-chain stumps are separated by pullulanase does the formation of iodine bands in the basic structure of the glycogen improve to a maximum of 500 nm. This maximum corresponds to unramified sections of less than 30 glucoside residues [4, 2, 6].  相似文献   

16.
Three phytosterols were isolated from Musa spp. flowers for evaluating their capabilities in inhibiting glucosidase and amylase activities and glycation of protein and sugar. The three phytosterols were identified as β‐sitosterol (PS1), 31‐norcyclolaudenone (PS2) and (24R)‐4α, 14α, 4‐trimethyl‐5α‐cholesta‐8, 25(27)‐dien‐3β‐ol (PS3). IC50 values (the concentration of inhibiting 50% of enzyme activity) of PS1, PS2 and PS3 against α‐glucosidase were 283.67, 11.33 and 43.10 μg mL?1, respectively. For inhibition of α‐amylase, the IC50 values of PS1, PS2 and PS3 were 52.55, 76.25 and 532.02 μg mL?1, respectively. PS1 was an uncompetitive inhibitor against α‐amylase with Km at 5.51 μg mL?1, while PS2 and PS3 exhibited a mixed‐type inhibition with Km at 52.36 and 2.49 μg mL?1, respectively. PS1 and PS2 also significantly inhibited the formation of advanced glycation end products (AGEs) in a BSA–fructose model. The results suggest that banana flower could possess the capability in prevention of the diseases associated with abnormal blood sugar and AGEs levels, such as diabetes.  相似文献   

17.
The hydrolysis of α-, β- and γ-cyclodextrins by Aspergillus oryzae α-amylase was studied at pH 5.2 and 37°C. The kinetic parameters were determined and it was found that the V max value increased markedly in the order α-, β- and γ-cyclodextrin, but no significant difference was observed in the Km values. The qualitative and quantitative distribution of the hydrolysis products were determined by HPLC. In the case of γ-cyclodextrin the time course of the kinetic parameters was compared to the qualitative and quantitative distribution of the hydrolysis products.  相似文献   

18.
α-Actinin release and its degradation from myofibrils Z-line were studied in post mortem white dorsal muscle from bass and sea trout stored at 4°C and 10°C. Using α-actinin specific antibodies, we show that this protein is rapidly released within the first 24 h for the two species, and reaches a plateau within 4 days. Proteolysis take place very rapidly in bass muscle yielding 80 and 40 kDa fragments from α-actinin as major bands of proteolysis. Sea trout muscle is more resistant, and muscle stored at 4°C is not significantly α-actinin degraded even 10 days after death. In the case of sea trout muscle stored at 10°C, an increasing quantity of 80 and 40 kDa fragment can be observed after the third day. These results show that release and proteolysis of α-actinin are time- and temperature-dependent processes that take place at the early stages of fish storage. Furthermore, we observed that proteolysis of α-actinin seems to be dependent on fish species. In both species studied, the early release of α-actinin comes before the degradation of released molecules, and appears as a biphasic process throughout the disorganisation of post mortem muscle in fish cold-stored above 0°C.  相似文献   

19.
We constructed two mouse α-amylase secretion vectors for Kluyveromyces lactis using the well-characterized signal sequence of the pGKL 128 kDa killer precursor protein. Both PHO5 and PGK expression cassettes from Saccharomyces cerevisiae directed the expression of mouse α-amylase in YPD medium at a similar level of efficiency. K. lactis transformants secreted glycosylated and non-glycosylated α-amylase into the culture medium and both species were enzymatically active. The K. lactis/S. cerevisiae shuttle secretion vector pMI6 was constructed, and K. lactis MD2/1(pMI6) secreted about four-fold more α-amylase than S. cerevisiae YNN27 harboring the same plasmid, indicating that K. lactis is an efficient host cell for the secretion and production of recombinant proteins. © 1997 John Wiley & Sons, Ltd.  相似文献   

20.
α-Amylases produced in germinated barley and incubated de-embryonated barley kernels (cv. Bonanza), in the absence and presence of gibberellic acid (GA3), were analyzed qualitatively by polyacrylamide gel isoelectric focusing (PAG-IEF) and quantitatively by chromatofocusing. Identical patterns of α-amylase components were obtained for both germinated barley and incubated de-embryonated barley kernels at each germination/incubation stage, in the absence or presence of GA3. Total α-amylase increased rapidly in the germinating whole seed whereas in the incubating de-embryonated grain the α-amylase activity increase was much slower. Addition of exogenous GA3 did not induce production of higher levels of α-amylase in either the germinating whole or incubating de-embryonated barley kernel. Quantitative chromatofocusing analysis revealed that the proportion of α-amylase III to α-amylase II activity decreased linearly with germination time in the whole grain but remained constant in the incubating de-embryonated grain in the absence or presence of GA3. The major proportion of α-amylase activity in the germinating whole grain and incubating de-embryonated grain was synthesized in the form of α-amylase II components. However, α-amylase I represented a larger proportion of the total α-amylase activity produced in the incubating de-embryonated grain, as compared to the germinating whole seed in the absence or presence of GA3. These results suggest that embryo excision differentially affects production of α-amylase II as compared to α-amylase I.  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