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.
This study reports on the determination of the depolymerization kinetics of amylose, amylopectin, and soluble starch by Aspergillus oryzae alpha-amylase using flow-injection analysis with fluorescence detection and 2-p-toluidinylnaphthalene-6-sulfonate as the fluorescent probe. The experimental data points, corresponding to the evolution of the concentration of "detectable" substrate with depolymerization time, were fit to a single exponential decay curve in the case of amylose and to a double exponential decay curve in the cases of amylopectin and soluble starch. For all the assayed substrates, the determined depolymerization rates at time zero correlated well with the initial enzyme and substrate concentrations through the usual Michaelis-Menten hyperbola. Therefore, this methodology allows the determination of alpha-amylase activity using any of these substrates. For amylopectin and soluble starch, the value of the total depolymerization rate at any depolymerization time was the result of the additive contribution of two partial depolymerization rates. In contrast, the total depolymerization rate for amylose was always a single value. These results, in conjunction with the relative time evolution of the two partial depolymerization rates (for amylopectin and soluble starch), are in good agreement with a linear molecular structure for amylose, a "grape-like" cluster molecular structure for amylopectin, and an extensively degraded grape-like cluster structure for soluble starch.  相似文献   

2.
Complexes of amylose and amylopectins with multivalent metal salts   总被引:1,自引:0,他引:1  
Metal cations [Cu(II), Fe(III), Mn(II), and Ni(II)] are ligated by amylose as well as potato, and corn amylopectins as proven by electron paramagnetic resonance spectra and conductivity measurements. The hydroxyl groups of polysaccharides are the coordination sites. Isolated starch polysaccharides did not coordinate to metal ions so well as starch did. The resulting polycenter Werner complexes were mainly square planar species. The ligation of the central metal atoms resulted in a variation of the thermal stability, pathway, and rate of thermal decomposition of starch as proven by thermogravimetric (TG, DTG) and scanning differential calorimetric measurements. Frequently, amylose and potato amylopectin willingly formed clathrates in which the water molecules were caged. The mode of the coordination of the hydroxyl groups to the central metal atom controlled the clathrate formation from amylose and in the case of potato amylopectin metal atoms bound to the phosphoric acid moiety formed cage by coordination of the hydroxyl groups to them. Coordination to selected metal salts controls pathway and products of polysaccharide ligand thermolysis.  相似文献   

3.
The specificity of Bacillus stearothermophilus TRS40 neopullulanase toward amylose and amylopectin was analyzed. Although this neopullulanase completely hydrolyzed amylose to produce maltose as the main product, it scarcely hydrolyzed amylopectin. The molecular mass of amylopectin was decreased by only one order of magnitude, from approximately 10(8) to 10(7) Da. Furthermore, this neopullulanase selectively hydrolyzed amylose when starch was used as a substrate. This phenomenon, efficient hydrolysis of amylose but not amylopectin, was also observed with cyclomaltodextrinase from alkaliphilic Bacillus sp. strain A2-5a and maltogenic amylase from Bacillus licheniformis ATCC 27811. These three enzymes hydrolyzed cyclomaltodextrins and amylose much faster than pullulan. Other amylolytic enzymes, such as bacterial saccharifying alpha-amylase, bacterial liquefying alpha-amylase, beta-amylase, and neopullulanase from Bacillus megaterium, did not exhibit this distinct substrate specificity at all, i.e., the preference of amylose to amylopectin.  相似文献   

4.
The extracellular alpha-amylase (1,4-alpha-D-glucanglucanohydrolase; EC 3.2.1.1) from maltose-grown Streptococcus bovis JB1 was purified to apparent homogeneity by ion-exchange chromatography (Mono Q). The enzyme had an isoelectric point of 4.50 and an apparent molecular mass of 77,000 Da, as estimated by sodium dodecyl sulfate-polyacrylamide gel electrophoresis. The enzyme was rich in acidic and hydrophobic amino acids. The 15-amino-acid NH2-terminal sequence was 40% homologous with the Bacillus subtilis saccharifying alpha-amylase and 27% homologous with the Clostridium acetobutylicum alpha-amylase. alpha-Amylase activity on soluble starch was optimal at pH 5.0 to 6.0. The enzyme was relatively stable between pH 5.5 and 8.5 and at temperatures below 50 degrees C. When soluble potato starch was used as the substrate, the enzyme had a Km of 0.88 mg.ml-1 and a kcat of 2,510 mumol of reducing sugar.min-1.mg of protein-1. The enzyme exhibited neither pullulanase nor dextranase activity and was 40 to 70% as active on amylopectin as on amylose. The major end products of amylose hydrolysis were maltose, maltotriose, and maltotetraose.  相似文献   

5.
Biochemical characterization of a novel heat-stable alpha-amylase, produced by a thermophilic strain of Bacillus brevis, has been made. The pattern of the enzyme action on different substrates was studied. It was found that reducing groups were rapidly liberated from amylopectin, soluble and insoluble starch compared to amylose and glycogen. B. brevis alpha-amylase acted via endo-attack producing mainly maltopentaose during the first hour of hydrolysis. The enzyme showed high activity towards maltohexaose and maltoheptaose. The alpha-amylase from B. brevis had a neutral pI and was found to be a glycoprotein, containing 9.2% (by mass) neutral sugars. The enzyme protein possessed a unique high glycine content. Calcium or sodium ions in appropriate concentrations were required for enzyme thermostability.  相似文献   

6.
Amylose prepared from starch dispersed in 10M-urea, pH6.2, was found to be resistant to the action of beta-amylase and phosphorylase, though it was degraded by alpha-amylase. Amylose isolated by conventional methods was similarly refractory after urea treatment, and was hydrolysed by beta-amylase to the extent of 32-35%; it had no inhibitory effect towards beta-amylase. The physical and chemical properties of the modified amylose were in general comparable with those of normal amylose with a beta-amylolysis limit of 94-98%. Starch and amylopectin were unaffected by urea treatment, i.e. the presence of amylopectin protected amylose against changes induced in it by urea. It is speculated that urea treatment "freezes" amylose molecules in a conformation that renders non-reducing termini inaccessible to the active site of the exo-enzymes. Such changes may limit the degradative action of beta-amylase and phosphorylase.  相似文献   

7.
Two constructs derived from the alpha-amylase gene (amyA) of Lactobacillus amylovorus were expressed in Lactobacillus plantarum, and their expression products were purified, characterized, and compared. These products correspond to the complete (AmyA) and truncated (AmyADelta) forms of alpha-amylase; AmyADelta lacks the 66-kDa carboxyl-terminal direct-repeating-unit region. AmyA and AmyADelta exhibit similar amylase activities towards a range of soluble substrates (amylose, amylopectin and alpha-cyclodextrin, and soluble starch). The specific activities of the enzymes towards soluble starch are similar, but the K(M) and V(max) values of AmyADelta were slightly higher than those of AmyA, whereas the thermal stability of AmyADelta was lower than that of AmyA. In contrast to AmyA, AmyADelta is unable to bind to beta-cyclodextrin and is only weakly active towards glycogen. More striking is the fact that AmyADelta cannot bind or hydrolyze raw starch, demonstrating that the carboxyl-terminal repeating-unit domain of AmyA is required for raw-starch binding activity.  相似文献   

8.
Damaged starch characterisation by ultracentrifugation   总被引:1,自引:0,他引:1  
The relative molecular size distributions of a selection of starches (waxy maize, pea and maize) that had received differing amounts of damage from ball milling (as quantified by susceptibility to alpha-amylase) were compared using analytical ultracentrifugation. Starch samples were solubilised in 90% dimethyl sulfoxide, and relative size distributions were determined in terms of the apparent distribution of sedimentation coefficients g*(s) versus s(20,w). For comparison purposes, the sedimentation coefficients were normalised to standard conditions of density and viscosity of water at 20 degrees C, and measurements were made with a standard starch loading concentration of 8 mg/mL. The modal molecular size of the native unmilled alpha-glucans were found to be approximately 50S, 51S and 79S for the waxy maize, pea and maize amylopectin molecules, respectively, whilst the pea and maize amylose modal molecular sizes were approximately 14S and approximately 12S, respectively. As the amount of damaged starch increased, the amylopectin molecules were eventually fragmented, and several components appeared, with the smallest fractions approaching the sedimentation coefficient values of amylose. For the waxy maize starch, the 50S material (amylopectin) was gradually converted to 14S, and the degradation process included the appearance of 24S material. For the pea starch, the situation was more complicated than the waxy maize due to the presence of amylose. As the amylopectin molecules (51S) were depolymerised by damage within this starch, low-molecular-weight fragments added to the proportion of the amylose fraction (14S)--although tending towards the high-molecular-weight region of this fraction. As normal maize starch was progressively damaged, a greater number of fragments appeared to be generated compared to the other two starches. Here, the 79S amylopectin peak (native starch) was gradually converted into 61 and 46S material and eventually to 11S material with a molecular size comparable to amylose. Amylose did not appear to be degraded, implying that all the damage was focused on the amylopectin fraction in all three cases. Specific differences in the damage profiles for the pea and maize starches may reflect the effect of lipid-complexed amylose in the maize starch.  相似文献   

9.
Readily available polysaccharides, amylopectin, amylose, dextrin, and yeast mannan, were chemically phosphorylated using polyphosphoric acid in the presence of a tertiary amine, and the resultant phosphates were examined for their interferon-inducing activity in rabbits employing an assay system consisting of a primary culture of rabbit kidney cells and vesicular stomatitis virus. All the phosphates were shown to be active as interferon inducer, and, especially, the activity of those containing more than 2% phosphorus were quite strong. Interferons evoked by the above phosphates resembled those induced by bacterial endotoxin, e.g., the viral inhibiting activity was susceptible to heat treatment, low pH and tryptic digestion. Since all the parent polysaccharides showed no interferon-inducing activity, it is reasonable to assume that the active center of these inducers might reside or be due to the anionic phosphate groups.  相似文献   

10.
The barley protein limit dextrinase inhibitor (LDI), structurally related to the alpha-amylase/trypsin inhibitor family, is an inhibitor of the starch debranching enzyme limit dextrinase (LD). In order to investigate the function of LDI, and the consequences for starch metabolism of reduced LDI activity, transgenic barley plants designed to downregulate LDI by antisense were generated. Homozygous antisense lines with reduced LDI protein level and activity were analysed and found to have enhanced free LD activity in both developing and germinating grains. In addition the antisense lines showed unpredicted pleiotropic effects on numerous enzyme activities, for example, alpha- and beta-amylases and starch synthases. Analysis of the starch showed much reduced numbers of the small B-type starch granules, as well as reduced amylose relative to amylopectin levels and reduced total starch. The chain length distribution of the amylopectin was modified with less of the longer chains (>25 units) and enhanced number of medium chains (10-15 units). These results suggest an important role for LDI and LD during starch synthesis as well as during starch breakdown.  相似文献   

11.
The distribution of substituents in hydroxypropylated potato amylopectin starch (amylose deficient) modified in a slurry of granular starch (HPPAPg) or in a polymer 'solution' of dissolved starch (HPPAPs), was investigated. The molar substitution (MS) was determined by three different methods: proton nuclear magnetic resonance (1H NMR) spectroscopy, gas-liquid chromatography (GLC) with mass spectrometry, and a colourimetric method. The MS values obtained by 1H NMR spectroscopy were higher than those obtained by GLC-mass spectrometry analysis and colourimetry. The relative ratio of 2-, 3-, and 6-substitution, as well as un-, mono-, and disubstitution in the anhydroglucose unit (AGU) were determined by GLC-mass spectrometry analysis. Results obtained showed no significant difference in molar distribution of hydroxypropyl groups in the AGU between the two derivatives. For analysis of the distribution pattern along the polymer chain, the starch derivatives were hydrolysed by enzymes with different selectivities. Debranching of the polymers indicated that more substituents were located in close vicinity to branching points in HPPAPg than in HPPAPs. Simultaneous alpha-amylase and amyloglucosidase hydrolysis of HPPAPg liberated more unsubstituted glucose units than the hydrolysis of HPPAPs, indicating a more heterogeneous distribution of substituents in HPPAPg.  相似文献   

12.
A low glycemic index starch was developed by partial alpha-amylase treatment, and its fine structure responsible for slowly digestible and resistant properties was investigated. Different digestion rates were obtained for gelatinized, retrograded starch by varying the enzyme dosage and reaction time. Analysis by high performance size-exclusion chromatography (HPSEC) coupled with multiangle laser-light scattering indicated that the molecular weighs of amylopectin and amylose were reduced during the digestion, to less than 100 kDa. A debranched chain length study using high performance anion-exchange chromatography equipped with an amyloglucosidase reactor and a pulsed amperometric detector and HPSEC revealed that short chains of amylopectin and noncrystalline amylose were rapidly digested, while DPn 121 chains showed resistance, followed by DPn 46 chains. X-ray diffraction analysis revealed that the crystalline structure in the treated starches survived cooking. These starches not only have slowly digestible and resistant character, but also retain some branched structure for adequate functionality.  相似文献   

13.
Aspergillus foetidus ATCC 10254 was selected from nine starch-utilizing microorganisms for its high amylolytic activity. This mould produced high levels of extracellular alpha-amylase in rice starch medium and degraded the available starch efficiently. Optimal conditions for enzyme production on 2.0% rice medium included 28 degrees C, initial pH of 6.6, and supplementations with 0.02% NaNO2, 0.08% KH2PO4, and 0.08% corn steep liquor. Eleven-fold purification of the enzyme was obtained after ammonium sulphate and ethanol precipitations from spent medium. The molecular weight was estimated at 41 500. Optimum pH and temperature for enzyme activity were 5.0 and 45 degrees C. Michaelis-Menten constants were 1.14 mg/ml on amylopectin, 2.19 mg/ml on soluble starch and 7.65 mg/ml on amylose. Amylose produced substrate inhibition while glucose or maltose did not inhibit the enzyme. This alpha-amylase may be used as a saccharifying enzyme for rice starch. Aspergillus foetidus ATCC 10254 also presents a potential for treatment of starch-containing waste waters.  相似文献   

14.
This protocol describes a high through put colorimetric method that relies on the formation of a complex between iodine and chains of glucose molecules in starch. Iodine forms complexes with both amylose and long chains within amylopectin. After the addition of iodine to a starch sample, the maximum absorption of amylose and amylopectin occurs at 620 and 550 nm, respectively. The amylose/amylopectin ratio can be estimated from the ratio of the 620 and 550 nm absorbance values and comparing them to a standard curve in which specific known concentrations are plotted against absorption values. This high throughput, inexpensive method is reliable and reproducible, allowing the evaluation of large populations of potato clones.   相似文献   

15.
An alpha-amylase produced by Scytalidium thermophilum was purified using DEAE-cellulose and CM-cellulose ion exchange chromatography and Sepharose 6B gel filtration. The purified protein migrated as a single band in 6% PAGE and 7% SDS-PAGE. The estimated molecular mass was 36 kDa (SDS-PAGE) and 49 kDa (Sepharose 6B). Optima of pH and temperature were 6.0 and 60 degrees C, respectively. In the absence of substrate the purified alpha-amylase was stable for 1 h at 50 degrees C and had a half-life of 12 min at 60 degrees C, but was fully stable in the presence of starch. The enzyme was not activated by several metal ions tested, including Ca(2+) (up to 10 mM), but HgCl(2 )and CuCl(2) inhibited its activity. The alpha-amylase produced by S. thermophilum preferentially hydrolyzed starch, and to a lesser extent amylopectin, maltose, amylose and glycogen in that order. The products of starch hydrolysis (up to 6 h of reaction) analyzed by thin layer chromatography, showed oligosaccharides such as maltotrioses, maltotetraoses and maltopentaoses. Maltose and traces of glucose were formed only after 3 h of reaction. These results confirm the character of the enzyme studied to be an alpha-amylase (1,4-alpha-glucan glucanohydrolase).  相似文献   

16.
alpha-Amylase production was examined in the ruminal anaerobic fungus Neocallimastix frontalis. The enzyme was released mainly into the culture fluid and had temperature and pH optima of 55 degrees C and 5.5, respectively, and the apparent K(m) for starch was 0.8 mg ml. The products of alpha-amylase action were mainly maltotriose, maltotetraose, and longer-chain oligosaccharides. No activity of the enzyme was observed towards these compounds or pullulan, but activity on amylose was similar to starch. Evidence for the endo action of alpha-amylase was also obtained from experiments which showed that the reduction in iodine-staining capacity and release in reducing power by action on amylose was similar to that for commercial alpha-amylase. Activities of alpha-amylase up to 4.4 U ml (1 U represents 1 mumol of glucose equivalents released per min) were obtained for cultures grown on 2.5 mg of starch ml in shaken cultures. No growth occurred in unshaken cultures. With elevated concentrations of starch (>2.5 mg ml), alpha-amylase production declined and glucose accumulated in the cultures. Addition of glucose to cultures grown on low levels of starch, in which little glucose accumulated, suppressed alpha-amylase production, and in bisubstrate growth studies, active production of the enzyme only occurred during growth on starch after glucose had been preferentially utilized. When cellulose, cellobiose, glucose, xylan, and xylose were tested as growth substrates for the production of alpha-amylase (initial concentration, 2.5 mg ml), they were found to be less effective than starch, but maltose was almost as effective. The fungal alpha-amylase was found to be stable at 60 degrees C in the presence of low concentrations of starch (相似文献   

17.
To study the effect of reagent type on the distribution pattern of acetyl groups in acetylated cowpea starch, amylose and amylopectin populations were isolated from the starch granules after modification to a low degree of substitution (DS < 0.1) with acetic anhydride and vinyl acetate, respectively. Slowly reacting reagent vinyl acetate resulted in higher DS values for the amylopectin populations when compared to the rapidly reacting reagent acetic anhydride. The two reagents had similar effects on the acetylation level of amylose, suggesting that the amorphous regions of granules were easily accessible for both reagents. The acetyl substitution patterns were analyzed by enzymatic degradation followed by characterization of the obtained fragments using chromatographic and mass spectrometric techniques. The distributions of acetyl groups along the amylose and amylopectin chains were more clustered for modification with vinyl acetate as compared with modification with acetic anhydride. Between the two acetylation types, pronounced differences in the acetyl substitution patterns were observed for the large fragments obtained after -amylase digestion; only slight differences were exhibited for the small fragments obtained by exhaustive enzymatic digestion of amylose and amylopectin populations.  相似文献   

18.
Common corn starch was modified in 0.56 M sodium sulfate solution and in 0.31 M potassium citrate solution. It was found that about 1.8 times the amount of reagent (propylene oxide) was needed to get a same molar substitution (MS) when potassium citrate was used. Hydroxypropylated starches were fractionated on a size-exclusion column to separate amylose from amylopectin, and MS values of the whole starch, the amylose, and the amylopectin were determined. In all preparations, amylose was derivatized to a greater extent than was amylopectin. The data indicate that, with common corn starch: (1) the greater the overall derivatization, the greater was the preference for derivatization of amylose; and (2) the preference for amylose derivatization was greater for corn starch modified in potassium citrate solution than in sodium sulfate solution when the MS values for the two preparations were essentially the same.  相似文献   

19.
Tang H  Hills BP 《Biomacromolecules》2003,4(5):1269-1276
To investigate the domain structure and dynamics of polysaccharides in the native starch granules, a variety of high resolution, solid-state (13)C NMR techniques have been applied to all three (A-, B-, and C-) types of starch with different water content. Both single-pulse-excitation magic-angle-spinning (SPEMAS) and cross-polarization-magic-angle-spinning (CPMAS) methods have been employed together with the PRISE (proton relaxation induced spectral-editing) techniques to distinguish polysaccharide fractions in different domains and having distinct dynamics. It has been found that, for all three types of dry starch granules, there are two sets of NMR signals corresponding to two distinct ordered polysaccharides. Hydration leads to substantial mobilization of the polysaccharides in the amorphous regions, but no fundamental changes in the rigidity of the polysaccharides in the crystalline (double) helices. Full hydration also leads to limited mobility changes to the polysaccharides in the amorphous lamellae (branching zone) within the amylopectin clusters and in the gaps between the arrays of the amylopectin clusters. Under magic-angle spinning, proton relaxation-time measurements showed a single component for T(1), two components for T(1rho), and three components for T(2). PRISE experiments permitted the neat separation of the (13)C resonances of polysaccharides in the crystalline lamellae from those in the amorphous lamellae and the amylose in the gaps between amylopectin clusters. It has been found that the long (1)H T(1rho) component ( approximately 30 ms) is associated with polysaccharides in the crystalline lamellae in the form of double helices, whereas the short T(1rho) component (2-4 ms) is associated with amylose in the gaps between amylopectin clusters. The short (1)H T(2) component ( approximately 14 micros) is associated with polysaccharides in the crystalline lamellae; the intermediate component (300-400 micros) is associated with polysaccharides in the amorphous lamellae and amylose in the gaps between amylopectin clusters. The long T(2) component is associated with both mobile starch protons and the residue water protons.  相似文献   

20.
He J  Liu J  Zhang G 《Biomacromolecules》2008,9(1):175-184
The mechanism and molecular structure of the slowly digestible waxy maize starch prepared by octenyl succinic anhydride (OSA) esterification and heat-moisture treatment were investigated. The in vitro Englyst test showed a proportion of 28.3% slowly digestible starch (SDS) when waxy maize starch was esterified with 3% OSA (starch weight based, and it is named OSA-starch), and a highest SDS content of 42.8% was obtained after OSA-starch (10% moisture) was further heated at 120 degrees C for 4 h (named HOSA-starch). The in vivo glycemic response of HOSA-starch, which showed a delayed appearance of blood glucose peak and a significant reduction (32.2%) of the peak glucose concentration, further confirmed its slow digestion property. Amylopectin debranching analysis revealed HOSA-starch had the highest resistance to debranching enzymes of isoamylase and pullulanase, and a simultaneous decrease of K m and V m (enzyme kinetics) was also shown when HOSA-starch was digested by either alpha-amylase or amyloglucosidase, indicating that the slow digestion of HOSA-starch resulted from an uncompetitive inhibition of enzyme activity during digestion. Size exclusion chromatography analysis of HOSA-starch showed fragmented amylopectin molecules with more nonreducing ends that are favorable for RS conversion to SDS by the action of amyloglucosidase in the Englyst test. Further solubility analysis indicates that the water-insolubility of HOSA-starch is caused by OSA-mediated cross-linking of amylopectin and the hydrophobic interaction between OSA-modified starch molecules. The water-insolubility of HOSA-starch would decrease its enzyme accessibility, and the digestion products with attached OSA molecules might also directly act as the uncompetitive inhibitor to reduce the enzyme activity leading to a slow digestion of HOSA-starch.  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