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.
AmyA, an alpha-amylase from the hyperthermophilic bacterium Thermotoga maritima, is able to hydrolyze internal alpha-1,4-glycosidic bonds in various alpha-glucans at 85 degrees C as the optimal temperature. Like other glycoside hydrolases, AmyA also catalyzes transglycosylation reactions, particularly when oligosaccharides are used as substrates. It was found that when methanol or butanol was used as the nucleophile instead of water, AmyA was able to catalyze alcoholysis reactions. This capability has been evaluated in the past for some alpha-amylases, with the finding that only the saccharifying fungal amylases from Aspergillus niger and from Aspergillus oryzae present measurable alcoholysis activity (R. I. Santamaria, G. Del Rio, G. Saab, M. E. Rodriguez, X. Soberon, and A. Lopez, FEBS Lett. 452:346-350, 1999). In the present work, we found that AmyA generates larger quantities of alkyl glycosides than any amylase reported so far. In order to increase the alcoholytic activity observed in AmyA, several residues were identified and mutated based on previous analogous positions in amylases, defining the polarity and geometry of the active site. Replacement of residue His222 by glutamine generated an increase in the alkyl glucoside yield as a consequence of a higher alcoholysis/hydrolysis ratio. The same change in specificity was observed for the mutants H222E and H222D, but instability of these mutants toward alcohols decreased the yield of alkyl glucoside.  相似文献   

2.
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   

3.
Two constructs derived from the α-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 (AmyAΔ) forms of α-amylase; AmyAΔ lacks the 66-kDa carboxyl-terminal direct-repeating-unit region. AmyA and AmyAΔ exhibit similar amylase activities towards a range of soluble substrates (amylose, amylopectin and α-cyclodextrin, and soluble starch). The specific activities of the enzymes towards soluble starch are similar, but the KM and Vmax values of AmyAΔ were slightly higher than those of AmyA, whereas the thermal stability of AmyAΔ was lower than that of AmyA. In contrast to AmyA, AmyAΔ is unable to bind to β-cyclodextrin and is only weakly active towards glycogen. More striking is the fact that AmyAΔ 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   

4.
We previously reported that Aspergillus oryzae strain S2 had produced two α-amylase isoforms named AmyA and AmyB. The apparent molecular masses revealed by SDS-PAGE were 50 and 42 kDa, respectively. Yet AmyB has a higher catalytic efficiency. Based on a monitoring study of the α-amylase production in both the presence and absence of different protease inhibitors, a chymotrypsin proteolysis process was detected in vivo generating AmyB. A. oryzae S2 α-amylase gene was amplified, cloned and sequenced. The sequence analysis revealed nine exons, eight introns and an encoding open reading frame of 1500 bp corresponding to AmyA isoform. The amino-acid sequence analysis revealed aY371 potential chymotrypsin cleaving site, likely to be the AmyB C-Terminal end and two other potential sites at Y359, and F379. A zymogram with a high acrylamide concentration was used. It highlighted two other closed apparent molecular mass α-amylases termed AmyB1 and AmyB2 reaching40 kDa and 43 kDa. These isoforms could be possibly generated fromY359, and F379secondary cut, respectively. The molecular modeling study showed that AmyB preserved the (β/α)8 barrel domain and the domain B but lacked the C-terminal domain C. The contact map analysis and the docking studies strongly suggested a higher activity and substrate binding affinity for AmyB than AmyA which was previously experimentally exhibited. This could be explained by the easy catalytic cleft accessibility.  相似文献   

5.
Alpha amylase family is generally defined as a group of enzymes that can hydrolyse and transglycosylase α-(1, 4) or α-(1, 6) glycosidic bonds along with the preservation of anomeric configuration. For the comparative analysis of alpha amylase family, nucleotide sequences of seven thermo stable organisms of Kingdom Archea i.e. Pyrococcus furiosus (100-105°C), Kingdom Prokaryotes i.e. Bacillus licheniformis (90-95°C), Geobacillus stearothermophilus (75°C), Bacillus amyloliquefaciens (72°C), Bacillus subtilis (70°C) and Bacillus KSM K38 (55°C) and Eukaryotes i.e. Aspergillus oryzae (60°C) were selected from NCBI. Primary structure composition analysis and Conserved sequence analysis were conducted through Bio Edit tools. Results from BioEdit shown only three conserved regions of base pairs and least similarity in MSA of the above mentioned alpha amylases. In Mega 5.1 Phylogeny of thermo stable alpha amylases of Kingdom Archea, Prokaryotes and Eukaryote was handled by Neighbor-Joining (NJ) algorithm. Mega 5.1 phylogenetic results suggested that alpha amylases of thermo stable organisms i.e. Pyrococcus furiosus (100-105°C), Bacillus licheniformis (90-95°C), Geobacillus stearothermophilus (75°C) and Bacillus amyloliquefaciens (72°C) are more distantly related as compared to less thermo stable organisms. By keeping in mind the characteristics of most thermo stable alpha amylases novel and improved features can be introduced in less thermo stable alpha amylases so that they become more thermo tolerant and productive for industry.  相似文献   

6.
Alkyl glycosides were synthesised from starch and alcohols using Aspergillus oryzae α-amylase as catalyst. In the degradation of starch by α-amylase, the alcohols competed with water as glycosyl acceptors. In the reaction with methanol, methyl maltoside and methyl maltotrioside were the main alcoholysis products. Conversion of 45 g/l starch in 30% methanol resulted in a product mixture containing 26 mM maltooligosaccharides and 3.6 mM methyl glycosides. With ethanol, propanol and butanol, alkyl maltosides and alkyl maltotetraosides were detected, and with benzyl alcohol, benzyl glycosides having two, three or five glucose units were formed. No alcoholysis reaction occurred with hexanol or octanol. In conclusion, α-amylase is promising for the one-step synthesis of alkyl glycosides having more than one monosaccharide unit, which are difficult to synthesise in other ways.  相似文献   

7.
Here, we established a system for displaying heterologous protein to the C terminus of the peptidoglycan-binding domain (cA domain) of AcmA (a major autolysin from Lactococcus lactis). Western blot and flow cytometric analyses revealed that the fusion proteins (cA-AmyA) of the cA domain and α-amylase from Streptococcus bovis 148 (AmyA) are efficiently expressed and successfully displayed on the surfaces of L. lactis cells. AmyA was also displayed on the cell surface while retaining its activity. Moreover, with an increase in the number of cA domains, the quantity of cA-AmyA fusion proteins displayed on the cell surface increased. When three repeats of the cA domain were used as an anchor protein, 82% of α-amylase activity was detected on the cells. The raw starch-degrading activity of AmyA was significantly higher when AmyA was fused to the C terminus of the cA domain than when it was fused to the N terminus. In addition, cA-AmyA fusion proteins were successfully displayed on the cell surfaces of Lactobacillus plantarum and Lactobacillus casei.  相似文献   

8.
In Saccharomyces cerevisiae, meiosis and spore formation as well as mating are controlled by mating-type genes. Diploids heterozygous for mating type (aα) can sporulate but cannot mate; homozygous aa and αα diploids can mate, but cannot sporulate. From an αα diploid parental strain, we have isolated mutants which have gained the ability to sporulate. Those mutants which continue to mate as αα cells have been designated CSP (control of sporulation). Upon sporulation, CSP mutants yield asci containing 4α spores. The mutant gene which allows αα cells to sporulate is unlinked to the mating-type locus and also acts to permit sporulation in aa diploid cells. Segregation data from crosses between mutant αα and wild-type aa diploids and vice versa indicate (for all but one mutant) that the mutation which allows constitutive sporulation (CSP) is dominant over the wild-type allele. Some of the CSP mutants are temperature-sensitive, sporulating at 32°, but not at 23°. In addition to CSP mutants, our mutagenesis and screening procedure led to the isolation of mutants which sporulate by virtue of a change in the mating-type locus itself, resulting in loss of ability to mate.  相似文献   

9.
D222G/N substitutions in A(H1N1)pdm09 hemagglutinin may be associated with increased binding of viruses causing low respiratory tract infections and human pathogenesis. We assessed the impact of such substitutions on the balance between hemagglutinin binding and neuraminidase cleavage, viral growth and in vivo virulence.Seven viruses with differing polymorphisms at codon 222 (2 with D, 3 G, 1 N and 1 E) were isolated from patients and characterized with regards hemagglutinin binding affinity (Kd) to α-2,6 sialic acid (SAα-2,6) and SAα-2,3 and neuraminidase enzymatic properties (Km, Ki and Vmax). The hemagglutination assay was used to quantitatively assess the balance between hemagglutinin binding and neuraminidase cleavage. Viral growth properties were compared in vitro in MDCK-SIAT1 cells and in vivo in BALB/c mice. Compared with D222 variants, the binding affinity of G222 variants was greater for SAα-2,3 and lower for SAα-2,6, whereas that of both E222 and N222 variants was greater for both SAα-2,3 and SAα-2,6. Mean neuraminidase activity of D222 variants (16.0 nmol/h/106) was higher than that of G222 (1.7 nmol/h/106 viruses) and E/N222 variants (4.4 nmol/h/106 viruses). The hemagglutination assay demonstrated a deviation from functional balance by E222 and N222 variants that displayed strong hemagglutinin binding but weak neuraminidase activity. This deviation impaired viral growth in MDCK-SIAT1 cells but not infectivity in mice. All strains but one exhibited low infectious dose in mice (MID50) and replicated to high titers in the lung; this D222 strain exhibited a ten-fold higher MID50 and replicated to low titers. Hemagglutinin-neuraminidase balance status had a greater impact on viral replication than hemagglutinin affinity strength, at least in vitro, thus emphasizing the importance of an optimal balance for influenza virus fitness. The mouse model is effective in assessing binding to SAα-2,3 but cannot differentiate SAα-2,3- from SAα-2,6- preference, nor estimate the hemagglutinin-neuraminidase balance in A(H1N1)pdm09 strains.  相似文献   

10.
The triethylammonium QX-314 and the trimethylammonium QX-222 are lidocaine derivatives that act as open-channel blockers of the acetylcholine (ACh) receptor. When bound, these blockers should occlude some of the residues lining the channel. Eight residues in the second membrane-spanning segment (M2) of the mouse-muscle α subunit were mutated one at a time to cysteine and expressed together with wild-type β, γ, and δ subunits in Xenopus oocytes. The rate constant for the reaction of each substituted cysteine with 2-aminoethyl methanethiosulfonate (MTSEA) was determined from the time course of the irreversible effect of MTSEA on the ACh-induced current. The reactions were carried out in the presence and absence of ACh and in the presence and absence of QX-314 and QX-222. These blockers had no effect on the reactions in the absence of ACh. In the presence of ACh, both blockers retarded the reaction of extracellularly applied MTSEA with cysteine substituted for residues from αVal255, one third of the distance in from the extracellular end of M2, to αGlu241, flanking the intracellular end of M2, but not with cysteine substituted for αLeu258 or αGlu262, at the extracellular end of M2. The reactions of MTSEA with cysteines substituted for αLeu258 and αGlu262 were considerably faster in the presence of ACh than in its absence. That QX-314 and QX-222 did not protect αL258C and αE262C against reaction with MTSEA in the presence of ACh implies that protection of the other residues was due to occlusion of the channel and not to the promotion of a less reactive state from a remote site. Given the 12-Å overall length of the blockers and the α-helical conformation of M2 in the open state, the binding site for both blockers extends from αVal255 down to αSer248.  相似文献   

11.
The identification of hereditary familial Alzheimer disease (FAD) mutations in the amyloid precursor protein (APP) and presenilin-1 (PS1) corroborated the causative role of amyloid-β peptides with 42 amino acid residues (Aβ42) in the pathogenesis of AD. Although most FAD mutations are known to increase Aβ42 levels, mutations within the APP GxxxG motif are known to lower Aβ42 levels by attenuating transmembrane sequence dimerization. Here, we show that aberrant Aβ42 levels of FAD mutations can be rescued by GxxxG mutations. The combination of the APP-GxxxG mutation G33A with APP-FAD mutations yielded a constant 60% decrease of Aβ42 levels and a concomitant 3-fold increase of Aβ38 levels compared with the Gly33 wild-type as determined by ELISA. In the presence of PS1-FAD mutations, the effects of G33A were attenuated, apparently attributable to a different mechanism of PS1-FAD mutants compared with APP-FAD mutants. Our results contribute to a general understanding of the mechanism how APP is processed by the γ-secretase module and strongly emphasize the potential of the GxxxG motif in the prevention of sporadic AD as well as FAD.  相似文献   

12.
Formation of the spore photoproduct (SP) (5-thyminyl-5,6-dihydrothymine) in DNA of dormant spores of Bacillus subtilis upon UV irradiation is due to binding of α/β-type small, acid-soluble proteins (SASP). However, the yield of SP as a function of UV fluence is ~15-fold higher in spores than in an α/β-type-SASP-DNA complex in vitro. The yield of SP as a function of UV fluence in forespore DNA from mutants which make α/β-type SASP but not dipicolinic acid (DPA) was 10 to 20 times lower than that in dormant spores. Furthermore, the yield of SP as a function of UV fluence in an α/β-type-SASP-DNA complex in vitro was increased sixfold by DPA. These data provide further support for the idea that the high DPA level in dormant spores increases the yield of SP as a function of UV fluence and thereby sensitizes spores to UV.  相似文献   

13.
A maltogenic amylase (MAG1) from alkaliphilic Bacillus lehensis G1 was cloned, expressed in Escherichia coli, purified and characterised for its hydrolysis and transglycosylation properties. The enzyme exhibited high stability at pH values from 7.0 to 10.0. The hydrolysis of β-cyclodextrin (β-CD) produced malto-oligosaccharides of various lengths. In addition to hydrolysis, MAG1 also demonstrated transglycosylation activity for the synthesis of longer malto-oligosaccharides. The thermodynamic equilibrium of the multiple reactions was shifted towards synthesis when the reaction conditions were optimised and the water activity was suppressed, which resulted in a yield of 38% transglycosylation products consisting of malto-oligosaccharides of various lengths. Thin layer chromatography and high-performance liquid chromatography analyses revealed the presence of malto-oligosaccharides with a higher degree of polymerisation than maltoheptaose, which has never been reported for other maltogenic amylases. The addition of organic solvents into the reaction further suppressed the water activity. The increase in the transglycosylation-to-hydrolysis ratio from 1.29 to 2.15 and the increased specificity toward maltopentaose production demonstrated the enhanced synthetic property of the enzyme. The high transglycosylation activity of maltogenic amylase offers a great advantage for synthesising malto-oligosaccharides and rare carbohydrates.  相似文献   

14.
Beta-Amylases from Alfalfa (Medicago sativa L.) Roots   总被引:8,自引:8,他引:0       下载免费PDF全文
Amylase was found in high activity (193 international units per milligram protein) in the tap root of alfalfa (Medicago sativa L. cv. Sonora). The activity was separated by gel filtration chromatography into two fractions with molecular weights of 65,700 (heavy amylase) and 41,700 (light amylase). Activity staining of electrophoretic gels indicated the presence of one isozyme in the heavy amylase fraction and two in the light amylase fraction. Three amylase isozymes with electrophoretic mobilities identical to those in the heavy and the light amylase fractions were the only amylases identified in crude root preparations. Both heavy and light amylases hydrolyzed amylopectin, soluble starch, and amylose but did not hydrolyze pullulan or β-limit dextrin. The ratio of viscosity change to reducing power production during starch hydrolysis was identical for both alfalfa amylase fractions and sweet potato β-amylase, while that of bacterial α-amylase was considerably higher. The identification of maltose and β-limit dextrin as hydrolytic end-products confirmed that these alfalfa root amylases are all β-amylases.  相似文献   

15.
Heterotrimeric G proteins, composed of α, β, and γ subunits, are activated by exchange of GDP for GTP on the Gα subunit. Canonically, Gα is stimulated by the guanine-nucleotide exchange factor (GEF) activity of ligand-bound G protein–coupled receptors. However, Gα subunits may also be activated in a noncanonical manner by members of the Ric-8 family, cytoplasmic proteins that also act as GEFs for Gα subunits. We used a signaling pathway active during Drosophila gastrulation as a model system to study Ric-8/Gα interactions. A component of this pathway, the Drosophila12/13 subunit, Concertina (Cta), is necessary to trigger actomyosin contractility during gastrulation events. Ric-8 mutants exhibit similar gastrulation defects to Cta mutants. Here we use a novel tissue culture system to study a signaling pathway that controls cytoskeletal rearrangements necessary for cellular morphogenesis. We show that Ric-8 regulates this pathway through physical interaction with Cta and preferentially interacts with inactive Cta and directs its localization within the cell. We also use this system to conduct a structure–function analysis of Ric-8 and identify key residues required for both Cta interaction and cellular contractility.  相似文献   

16.
Three isoforms of α-glucosidase (EC 3.2.1.20) have been extracted from pea (Pisum sativum L.) seedlings and separated by DEAE-cellulose and CM-Sepharose chromatography. Two α-glucosidase isoforms (αG1 and αG2) were most active under acid conditions, and appeared to be apoplastic. A neutral form (αG3) was most active near pH 7, and was identified as a chloroplastic enzyme. Together, the activity of αG1 and αG2 in apoplastic preparations accounted for 21% of the total acid α-glucosidase activity recovered from pea stems. The vast majority (86%) of the apoplastic acid α-glucosidase activity was due to αG1. The apparent Km values for maltose of αG1 and αG2 were 0.3 and 1.3 millimolar, respectively. The apparent Km for maltose of αG3 was 33 millimolar. The respective native molecular weights of αG1, αG2, and αG3 were 125,000, 150,000, and 110,000.  相似文献   

17.
We have used gene disruption to isolate two talin (−/−) ES cell mutants that contain no intact talin. The undifferentiated cells (a) were unable to spread on gelatin or laminin and grew as rounded colonies, although they were able to spread on fibronectin (b) showed reduced adhesion to laminin, but not fibronectin (c) expressed much reduced levels of β1 integrin, although levels of α5 and αV were wild-type (d) were less polarized with increased membrane protrusions compared with a vinculin (−/−) ES cell mutant (e) were unable to assemble vinculin or paxillin-containing focal adhesions or actin stress fibers on fibronectin, whereas vinculin (−/−) ES cells were able to assemble talin-containing focal adhesions. Both talin (−/−) ES cell mutants formed embryoid bodies, but differentiation was restricted to two morphologically distinct cell types. Interestingly, these differentiated talin (−/−) ES cells were able to spread and form focal adhesion-like structures containing vinculin and paxillin on fibronectin. Moreover, the levels of the β1 integrin subunit were comparable to those in wild-type ES cells. We conclude that talin is essential for β1 integrin expression and focal adhesion assembly in undifferentiated ES cells, but that a subset of differentiated cells are talin independent for both characteristics.  相似文献   

18.
Interactions of the amyloid β-protein (Aβ) with neuronal cell membranes, leading to the disruption of membrane integrity, are considered to play a key role in the development of Alzheimer’s disease. Natural mutations in Aβ 42, such as the Arctic mutation (E22G) have been shown to increase Aβ 42 aggregation and neurotoxicity, leading to the early-onset of Alzheimer’s disease. A correlation between the propensity of Aβ 42 to form protofibrils and its effect on neuronal dysfunction and degeneration has been established. Using rational mutagenesis of the Aβ 42 peptide it was further revealed that the aggregation of different Aβ 42 mutants in lipid membranes results in a variety of polymorphic aggregates in a mutation dependent manner. The mutant peptides also have a variable ability to disrupt bilayer integrity. To further test the connection between Aβ 42 mutation and peptide–membrane interactions, we perform molecular dynamics simulations of membrane-inserted Aβ 42 variants (wild-type and E22G, D23G, E22G/D23G, K16M/K28M and K16M/E22G/D23G/K28M mutants) as β-sheet monomers and tetramers. The effects of charged residues on transmembrane Aβ 42 stability and membrane integrity are analyzed at atomistic level. We observe an increased stability for the E22G Aβ 42 peptide and a decreased stability for D23G compared to wild-type Aβ 42, while D23G has the largest membrane-disruptive effect. These results support the experimental observation that the altered toxicity arising from mutations in Aβ is not only a result of the altered aggregation propensity, but also originates from modified Aβ interactions with neuronal membranes.  相似文献   

19.
α-Galactosidases are broadly used in feed, food, chemical, pulp, and pharmaceutical industries. However, there lacks a satisfactory microbial cell factory that is able to produce α-galactosidases efficiently and cost-effectively to date, which prevents these important enzymes from greater application. In this study, the secretory expression of an Aspergillus niger α-galactosidase (AGA) in Pichia pastoris was systematically investigated. Through codon optimization, signal peptide replacement, comparative selection of host strain, and saturation mutagenesis of the P1’ residue of Kex2 protease cleavage site for efficient signal peptide removal, a mutant P. pastoris KM71H (Muts) strain of AGA-I with the specific P1’ site substitution (Glu to Ile) demonstrated remarkable extracellular α-galactosidase activity of 1299 U/ml upon a 72 h methanol induction in 2.0 L fermenter. The engineered yeast strain AGA-I demonstrated approximately 12-fold higher extracellular activity compared to the initial P. pastoris strain. To the best of our knowledge, this represents the highest yield and productivity of a secreted α-galactosidase in P. pastoris, thus holding great potential for industrial application.  相似文献   

20.
An examination of the carotene fractions extracted from Euglena gracilis Z and pressure-bleached Euglena mutants PR-1, PR-2, PR-3, and PR-4 revealed phytoene in mutants PR-1, PR-2, and PR-3. Photosynthetic E. gracilis Z cultured at different light intensities showed no detectable phytoene, nor was phytoene found in mutant PR-4. However, dark-cultured E. gracilis Z yielded readily assayable amounts of phytoene. With the exception of PR-4, in which no C40 carotenoids were detected, the following carotenes were identified in all from their mass spectra: phytoene, phytofluene, ζ-carotene, β-zeacarotene, and β-carotene. Of these, phytoene and β-zeacarotene had not previously been unequivocally identified in Euglena.  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