首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
The gene for a novel alpha-amylase, designated AmyC, from the hyperthermophilic bacterium Thermotoga maritima was cloned and heterologously overexpressed in Escherichia coli. The putative intracellular enzyme had no amino acid sequence similarity to glycoside hydrolase family (GHF) 13 alpha-amylases, yet the range of substrate hydrolysis and the product profile clearly define the protein as an alpha-amylase. Based on sequence similarity AmyC belongs to a subgroup within GHF 57. On the basis of amino acid sequence similarity, Glu185 and Asp349 could be identified as the catalytic residues of AmyC. Using a 60-min assay, the maximum hydrolytic activity of the purified enzyme, which was dithiothreitol dependent, was found to be at 90 degrees C. AmyC displayed a remarkably high pH optimum of pH 8.5 and an unusual sensitivity towards both ATP and EDTA.  相似文献   

2.
The cDNA coding for Penicillium purpurogenum α-galactosidase (αGal) was cloned and sequenced. The deduced amino acid sequence of the α-Gal cDNA showed that the mature enzyme consisted of 419 amino acid residues with a molecular mass of 46,334 Da. The derived amino acid sequence of the enzyme showed similarity to eukaryotic αGals from plants, animals, yeasts, and filamentous fungi. The highest similarity observed (57% identity) was to Trichoderma reesei AGLI. The cDNA was expressed in Saccharomyces cerevisiae under the control of the yeast GAL10 promoter. Almost all of the enzyme produced was secreted into the culture medium, and the expression level reached was approximately 0.2 g/liter. The recombinant enzyme purified to homogeneity was highly glycosylated, showed slightly higher specific activity, and exhibited properties almost identical to those of the native enzyme from P. purpurogenum in terms of the N-terminal amino acid sequence, thermoactivity, pH profile, and mode of action on galacto-oligosaccharides.  相似文献   

3.
Two α-amylase genes from the thermophilic alkaliphile Anaerobranca gottschalkii were cloned, and the corresponding enzymes, AmyA and AmyB, were investigated after purification of the recombinant proteins. Based on their amino acid sequences, AmyA is proposed to be a lipoprotein with extracellular localization and thus is exposed to the alkaline milieu, while AmyB apparently represents a cytoplasmic enzyme. The amino acid sequences of both enzymes bear high similarity to those of GHF13 proteins. The different cellular localizations of AmyA and AmyB are reflected in their physicochemical properties. The alkaline pH optimum (pH 8), as well as the broad pH range, of AmyA activity (more than 50% activity between pH 6 and pH 9.5) mirrors the conditions that are encountered by an extracellular enzyme exposed to the medium of A. gottschalkii, which grows between pH 6 and pH 10.5. AmyB, on the other hand, has a narrow pH range with a slightly acidic pH optimum at 6 to 6.5, which is presumably close to the pH in the cytoplasm. Also, the intracellular AmyB is less tolerant of high temperatures than the extracellular AmyA. While AmyA has a half-life of 48 h at 70°C, AmyB has a half-life of only about 10 min at that temperature, perhaps due to the lack of stabilizing constituents of the cytoplasm. AmyA and AmyB were very similar with respect to their substrate specificity profiles, clearly preferring amylose over amylopectin, pullulan, and glycogen. Both enzymes also hydrolyzed α-, β-, and γ-cyclodextrin. Very interestingly, AmyA, but not AmyB, displayed high transglycosylation activity on maltooligosaccharides and also had significant β-cyclodextrin glycosyltransferase (CGTase) activity. CGTase activity has not been reported for typical α-amylases before. The mechanism of cyclodextrin formation by AmyA is unknown.  相似文献   

4.
5.
Complete (Ba-L) and truncated (Ba-S) forms of α-amylases from Bacillus subtilis X-23 were purified, and the amino- and carboxyl-terminal amino acid sequences of Ba-L and Ba-S were determined. The amino acid sequence deduced from the nucleotide sequence of the α-amylase gene indicated that Ba-S was produced from Ba-L by truncation of the 186 amino acid residues at the carboxyl-terminal region. The results of genomic Southern analysis and Western analysis suggested that the two enzymes originated from the same α-amylase gene and that truncation of Ba-L to Ba-S occurred during the cultivation of B. subtilis X-23 cells. Although the primary structure of Ba-S was approximately 28% shorter than that of Ba-L, the two enzyme forms had the same enzymatic characteristics (molar catalytic activity, amylolytic pattern, transglycosylation ability, effect of pH on stability and activity, optimum temperature, and raw starch-binding ability), except that the thermal stability of Ba-S was higher than that of Ba-L. An analysis of the secondary structure as well as the predicted three-dimensional structure of Ba-S showed that Ba-S retained all of the necessary domains (domains A, B, and C) which were most likely to be required for functionality as α-amylase.  相似文献   

6.
1. β-Amylase obtained by acidic extraction of soya-bean flour was purified by ammonium sulphate precipitation, followed by chromatography on calcium phosphate, diethylaminoethylcellulose, Sephadex G-25 and carboxymethylcellulose. 2. The homogeneity of the pure enzyme was established by criteria such as ultracentrifugation and electrophoresis on paper and in polyacrylamide gel. 3. The pure enzyme had a nitrogen content of 16·3%, its extinction coefficient, E1%1cm., at 280mμ was 17·3 and its specific activity/mg. of enzyme was 880 amylase units. 4. The molecular weight of the pure enzyme was determined as 61700 and its isoelectric point was pH5·85. 5. Preliminary examinations indicated that glutamic acid formed the N-terminus and glycine the C-terminus. 6. The amino acid content of the pure enzyme was established, one molecule consisting of 617 amino acid residues. 7. The pH optimum for pure soya-bean β-amylase is in the range 5–6. Pretreatment of the enzyme at pH3–5 decreases enzyme activity, whereas at pH6–9 it is not affected.  相似文献   

7.
A novel liquefying α-amylase (LAMY) was found in cultures of an alkaliphilic Bacillus isolate, KSM-1378. The specific activity of purified LAMY was approximately 5,000 U mg of protein−1, a value two- to fivefold greater between pH 5 and 10 than that of an industrial, thermostable Bacillus licheniformis enzyme. The enzyme had a pH optimum of 8.0 to 8.5 and displayed maximum activity at 55°C. The molecular mass deduced from sodium dodecyl sulfate-polyacrylamide gel electrophoresis was approximately 53 kDa, and the apparent isoelectric point was around pH 9. This enzyme efficiently hydrolyzed various carbohydrates to yield maltotriose, maltopentaose, maltohexaose, and maltose as major end products after completion of the reaction. Maltooligosaccharides in the maltose-to-maltopentaose range were unhydrolyzable by the enzyme. The structural gene for LAMY contained a single open reading frame 1,548 bp in length, corresponding to 516 amino acids that included a signal peptide of 31 amino acids. The calculated molecular mass of the extracellular mature enzyme was 55,391 Da. LAMY exhibited relatively low amino acid identity to other liquefying amylases, such as the enzymes from B. licheniformis (68.9%), Bacillus amyloliquefaciens (66.7%), and Bacillus stearothermophilus (68.6%). The four conserved regions, designated I, II, III, and IV, and the putative catalytic triad were found in the deduced amino acid sequence of LAMY. Essentially, the sequence of LAMY was consistent with the tertiary structures of reported amylolytic enzymes, which are composed of domains A, B, and C and which include the well-known (α/β)8 barrel motif in domain A.  相似文献   

8.
Mutational experiments were carried out to decrease the protease productivity of Aspergillus ficum IFO 4320 by using N-methyl-N′-nitro-N-nitrosoguanidine. A protease-negative mutant, M-33, exhibited higher α-amylaseactivity than the parent strain under submerged culture at 30°C for 24 h. About 70% of the total α-amylase activity in the M-33 culture filtrate was adsorbed onto starch granules. The electrophoretically homogeneous preparation of raw-starch-adsorbable α-amylase (molecular weight, 88,000), acid stable at pH 2, showed intensive raw-starch-digesting activity, dissolving corn starch granules completely. The preparation also exhibited a high synergistic effect with glucoamylase I. A mutant, M-72, with higher protease activity produced a raw cornstarch-unadsorbable α-amylase. The purified enzyme (molecular weight, 54,000), acid unstable, showed no digesting activity on raw corn starch and a lower synergistic effect with glucoamylase I in the hydrolysis of raw corn starch. The fungal α-amylase was therefore divided into two types, a novel type of raw-starch-digesting enzyme and a conventional type of raw-starch-nondigesting enzyme.  相似文献   

9.
Sequencing upstream of the Streptococcus mutans gene for a CcpA gene homolog, regM, revealed an open reading frame, named amy, with homology to genes encoding α-amylases. The deduced amino acid sequence showed a strong similarity (60% amino acid identity) to the intracellular α-amylase of Streptococcus bovis and, in common with this enzyme, lacked a signal sequence. Amylase activity was found only in S. mutans cell extracts, with no activity detected in culture supernatants. Inactivation of amy by insertion of an antibiotic resistance marker confirmed that S. mutans has a single α-amylase activity. The amylase activity was induced by maltose but not by starch, and no acid was produced from starch. S. mutans can, however, transport limit dextrins and maltooligosaccharides generated by salivary amylase, but inactivation of amy did not affect growth on these substrates or acid production. The amylase digested the glycogen-like intracellular polysaccharide (IPS) purified from S. mutans, but the amy mutant was able to digest and produce acid from IPS; thus, amylase does not appear to be essential for IPS breakdown. However, when grown on excess maltose, the amy mutant produced nearly threefold the amount of IPS produced by the parent strain. The role of Amy has not been established, but Amy appears to be important in the accumulation of IPS in S. mutans grown on maltose.  相似文献   

10.
An inhibitor of malted barley (Hordeum vulgare cv Conquest) α-amylase II was purified 125-fold from a crude extract of barley kernels by (NH4)2SO4 fractionation, ion exchange chromatography on DEAE-Sephacel, and gel filtration on Bio-Gel P 60. The inhibitor was a protein with an approximate molecular weight of 20,000 daltons and an isoelectric point of 7.3. The protein was homogeneous, as assessed by sodium dodecyl sulfate-polyacrylamide gel electrophoresis. Amino acid analysis indicated the presence of about 9 half-cystine residues per mole. The neutral isoelectric point of the inhibitor suggested that some of the apparently acidic residues (glutamic and aspartic) existed in the amide form. The first twenty N-terminal amino acids were sequenced. Some homology appeared to exist between the α-amylase II inhibitor and trypsin inhibitor from barley. Complex formation between α-amylase II and the inhibitor was detected by the appearance of a new molecular weight species after gel filtration on Bio-Gel P 100. Enzyme and inhibitor had to be preincubated for 5 min, prior to assaying for enzyme activity before maximum inhibition was attained. Inhibition increased at higher pH values. At pH 5.5, an approximately 1100 molar excess of inhibitor over α-amylase II produced 40% inhibition, whereas, at pH 8.0, a 1:1 molar ratio of inhibitor to enzyme produced the same degree of inhibition.  相似文献   

11.
A β-amylase-overproducing mutant of Clostridium thermosulfurogenes was grown in continuous culture on soluble starch to produce thermostable β-amylase. Enzyme productivity was reasonably stable over periods of weeks to months. The pH and temperature optima for β-amylase production were pH 6.0 and 60°C, respectively. Enzyme concentration was maximized by increasing biomass concentration by using high substrate concentrations and by maintaining a low growth rate. β-Amylase concentration reached 90 U ml−1 at a dilution rate of 0.07 h−1 in a 3% starch medium. A further increase in enzyme activity levels was limited by acetic acid inhibition of growth and low β-amylase productivity at low growth rates.  相似文献   

12.
The most abundant β-amylase (EC 3.2.1.2) in pea (Pisum sativum L.) was purified greater than 880-fold from epicotyls of etiolated germinating seedlings by anion exchange and gel filtration chromatography, glycogen precipitation, and preparative electrophoresis. The electrophoretic mobility and relative abundance of this β-amylase are the same as that of an exoamylase previously reported to be primarily vacuolar. The enzyme was determined to be a β-amylase by end product analysis and by its inability to hydrolyze β-limit dextrin and to release dye from starch azure. Pea β-amylase is an approximate 55 to 57 kilodalton monomer with a pl of 4.35, a pH optimum of 6.0 (soluble starch substrate), an Arrhenius energy of activation of 6.28 kilocalories per mole, and a Km of 1.67 milligrams per milliliter (soluble starch). The enzyme is strongly inhibited by heavy metals, p-chloromer-curiphenylsulfonic acid and N-ethylmaleimide, but much less strongly by iodoacetamide and iodoacetic acid, indicating cysteinyl sulfhydryls are not directly involved in catalysis. Pea β-amylase is competitively inhibited by its end product, maltose, with a Ki of 11.5 millimolar. The enzyme is partially inhibited by Schardinger maltodextrins, with α-cyclohexaamylose being a stronger inhibitor than β-cycloheptaamylose. Moderately branched glucans (e.g. amylopectin) were better substrates for pea β-amylase than less branched or non-branched (amyloses) or highly branched (glycogens) glucans. The enzyme failed to hydrolyze native starch grains from pea and glucans smaller than maltotetraose. The mechanism of pea β-amylase is the multichain type. Possible roles of pea β-amylase in cellular glucan metabolism are discussed.  相似文献   

13.
The Neisseria polysaccharea gene encoding amylosucrase was subcloned and expressed in Escherichia coli. Sequencing revealed that the deduced amino acid sequence differs significantly from that previously published. Comparison of the sequence with that of enzymes of the α-amylase family predicted a (β/α)8-barrel domain. Six of the eight highly conserved regions in amylolytic enzymes are present in amylosucrase. Among them, four constitute the active site in α-amylases. These sites were also conserved in the sequence of glucosyltransferases and dextransucrases. Nevertheless, the evolutionary tree does not show strong homology between them. The amylosucrase was purified by affinity chromatography between fusion protein glutathione S-transferase–amylosucrase and glutathione-Sepharose 4B. The pure enzyme linearly elongated some branched chains of glycogen, to an average degree of polymerization of 75.  相似文献   

14.
The biosynthetic mechanism of α-amylase synthesis in germinating rice (Oryza sativa L. cv. Kimmazé) seeds has been studied both in vitro and in vivo. Special attention has been focused on the glycosylation of the enzyme molecule. Tunicamycin was found to inhibit glycosylation of α-amylase by 98% without significant inhibition of enzyme secretion. The inhibitory effect exerted by the antibiotic on glycosylation did not significantly alter enzyme activity.

In an in vitro system using poly-(A) RNA isolated from rice scutellum and the reticulocyte lysate translation system, a precursor form of α-amylase (precursor I) is formed. Inhibition of glycosylation by Tunicamycin allowed detection of a nonglycosylated precursor (II) of α-amylase. The molecular weight of the nonglycosylated precursor II produced in the presence of Tunicamycin was 2,900 daltons less than that of the mature form of α-amylase (44,000) produced in the absence of Tunicamycin, and 1,800 daltons less than the in vitro synthesized molecule.

The inhibition of glycosylation by Tunicamycin as well as in vitro translation helped clarify the heterogeneity of α-amylase isozymes. Isoelectrofocusing (pH 4-6) of the products, zymograms, and fluorography were employed on the separated isozyme components. The mature and Tunicamycin-treated nonglycosylated forms of α-amylase were found to consist of three isozymes. The in vitro translated precursor forms of α-amylase consisted of four multiple components. These results indicate that heterogeneity of α-amylase isozymes is not due to glycosylation of the enzyme protein but likely to differences in the primary structure of the protein moiety, which altogether support that rice α-amylase isozymes are encoded by multiple genes.

  相似文献   

15.
Clostridium thermosulfurogenes EM1 formed blebs, i.e., protrusions still in contact with the cytoplasmic membrane, that originated from the cytoplasmic membrane during growth in batch culture and continuous culture. They could be observed squeezed between the cell wall and cytoplasmic membrane in cells with seemingly intact wall layers (surface layer and peptidoglycan layer) as well as in cells with wall layers in different states of degradation caused by phosphate limitation or high dilution rates. Blebs were found to turn into membrane vesicles by constriction in cases when the cell wall was heavily degraded. Bleb and vesicle formation was also observed in the absence of substrates that induce α-amylase and pullulanase synthesis. No correlations existed between bleb formation and the presence of active enzyme. Similar blebs could also be observed in a number of other gram-positive bacteria not producing these enzymes, but they were not observed in gram-negative bacteria. For immunoelectron-microscopic localization of α-amylase and pullulanase in C. thermosulfurogenes EM1, two different antisera were applied. One was raised against the enzymes isolated from the culture fluid; the other was produced against a peptide synthesized, as a defined epitope, in analogy to the N-terminal amino acid sequence (21 amino acids) of the native extracellular α-amylase. By using these antisera, α-amylase and pullulanase were localized at the cell periphery in samples taken from continuous culture or batch culture. In samples prepared for electron microscopy by freeze substitution followed by ultrathin sectioning, blebs could be seen, and the immunolabel pinpointing α-amylase enzyme particles was seen not only randomly distributed in the cell periphery, but also lining the surface of the cytoplasmic membrane and the blebs. Cells exhibiting high or virtually no enzyme activity were labeled similarly with both antisera. This finding strongly suggests that α-amylase and pullulanase may occur in both active and inactive forms, depending on growth conditions.  相似文献   

16.
The β-galactosidase from the Antarctic gram-negative bacterium Pseudoalteromonas haloplanktis TAE 79 was purified to homogeneity. The nucleotide sequence and the NH2-terminal amino acid sequence of the purified enzyme indicate that the β-galactosidase subunit is composed of 1,038 amino acids with a calculated Mr of 118,068. This β-galactosidase shares structural properties with Escherichia coli β-galactosidase (comparable subunit mass, 51% amino sequence identity, conservation of amino acid residues involved in catalysis, similar optimal pH value, and requirement for divalent metal ions) but is characterized by a higher catalytic efficiency on synthetic and natural substrates and by a shift of apparent optimum activity toward low temperatures and lower thermal stability. The enzyme also differs by a higher pI (7.8) and by specific thermodynamic activation parameters. P. haloplanktis β-galactosidase was expressed in E. coli, and the recombinant enzyme displays properties identical to those of the wild-type enzyme. Heat-induced unfolding monitored by intrinsic fluorescence spectroscopy showed lower melting point values for both P. haloplanktis wild-type and recombinant β-galactosidase compared to the mesophilic enzyme. Assays of lactose hydrolysis in milk demonstrate that P. haloplanktis β-galactosidase can outperform the current commercial β-galactosidase from Kluyveromyces marxianus var. lactis, suggesting that the cold-adapted β-galactosidase could be used to hydrolyze lactose in dairy products processed in refrigerated plants.  相似文献   

17.
The coniferyl aldehyde dehydrogenase (CALDH) of Pseudomonas sp. strain HR199 (DSM7063), which catalyzes the NAD+-dependent oxidation of coniferyl aldehyde to ferulic acid and which is induced during growth with eugenol as the carbon source, was purified and characterized. The native protein exhibited an apparent molecular mass of 86,000 ± 5,000 Da, and the subunit mass was 49.5 ± 2.5 kDa, indicating an α2 structure of the native enzyme. The optimal oxidation of coniferyl aldehyde to ferulic acid was obtained at a pH of 8.8 and a temperature of 26°C. The Km values for coniferyl aldehyde and NAD+ were about 7 to 12 μM and 334 μM, respectively. The enzyme also accepted other aromatic aldehydes as substrates, whereas aliphatic aldehydes were not accepted. The NH2-terminal amino acid sequence of CALDH was determined in order to clone the encoding gene (calB). The corresponding nucleotide sequence was localized on a 9.4-kbp EcoRI fragment (E94), which was subcloned from a Pseudomonas sp. strain HR199 genomic library in the cosmid pVK100. The partial sequencing of this fragment revealed an open reading frame of 1,446 bp encoding a protein with a relative molecular weight of 51,822. The deduced amino acid sequence, which is reported for the first time for a structural gene of a CALDH, exhibited up to 38.5% amino acid identity (60% similarity) to NAD+-dependent aldehyde dehydrogenases from different sources.  相似文献   

18.
Ho TH  Abroms J  Varner JE 《Plant physiology》1982,69(5):1128-1131
A large portion of the gibberellic acid (GA3)-induced α-amylase in isolated aleurone layers is transported into the incubation medium. In the presence of GA3 and ethylene, an even larger portion of the enzyme is found in the medium. Employing an acid washing technique developed by Varner and Mense (Plant Physiol 1972 49:187-189), it was observed that ethylene significantly reduces the amount of α-amylase trapped by the thick cell walls of aleurone layers. However, the amount of enzyme remaining in the cell (within the boundary of plasma membrane) is not affected by ethylene. Ethylene has no observable effect on membrane formation as measured by the incorporation of [32P]orthophosphate into phospholipids. Because of these observations it is suggested that ethylene enhances the release of α-amylase, i.e. transport of α-amylase across cell walls, but not the secretion of α-amylase, i.e. transport of α-amylase past the barrier of plasma membrane. The possible mechanism of this ethylene effect is discussed.  相似文献   

19.
Scutellar tissues were dissected from germinating rice seeds and the incorporation of [35S]methionine into the α-amylase molecule was examined by in vivo and in vitro assay systems. Immunoprecipitation with monospecific anti-α-amylase immunoglobulin G raised against the purified enzyme preparation and sodium dodecyl sulfate-polyacrylamide gel electrophoresis and fluorography were used to identify α-amylase and its possible precursor molecule. Using freshly prepared scutellar tissues, it was demonstrated that α-amylase is synthesized de novo in the scutellar epithelium and secreted into endosperm. The synthesis of α-amylase directed by the polyadenylic acid-containing ribonucleic acid isolated from the scutellar tissues was also established using the translation system of either wheat germ extract or reticulocyte lysate. The immunoprecipitable product obtained in the in vitro translation system was smaller in molecular weight than that synthesized in vivo on the basis of mobilities in sodium dodecyl sulfate-polyacrylamide gel electrophoresis. Results are discussed in relation to the processing of the nascent polypeptide precursor of the enzyme molecule and the introduction of the oligosaccharide chain to the cleaved polypeptide to make up the mature form of α-amylase.  相似文献   

20.
β-Xylosidase secreted by the shoyu koji mold, Aspergillus oryzae, is the key enzyme responsible for browning of soy sauce. To investigate the role of β-xylosidase in the brown color formation, a major β-xylosidase, XylA, and its encoding gene were characterized. β-Xylosidase XylA was purified to homogeneity from culture filtrates of A. oryzae KBN616. The optimum pH and temperature of the enzyme were found to be 4.0 and 60°C, respectively, and the molecular mass was estimated to be 110 kDa based on sodium dodecyl sulfate-polyacrylamide gel electrophoresis. The xylA gene comprises 2,397 bp with no introns and encodes a protein consisting of 798 amino acids (86,475 Da) with 14 potential N-glycosylation sites. The deduced amino acid sequence shows high similarity to Aspergillus nidulans XlnD (70%), Aspergillus niger XlnD (64%), and Trichoderma reesei BxII (63%). The xylA gene was overexpressed under control of the strong and constitutive A. oryzae TEF1 promoter. One of the A. oryzae transformants produced approximately 13 times more of the enzyme than did the host strain. The partial-length antisense xylA gene expressed under control of the A. oryzae TEF1 promoter decreased the β-xylosidase level in A. oryzae to about 20% of that of the host strain.  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