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.
Dextran glucosidase from Streptococcus mutans (SMDG), an exo-type glucosidase of glycoside hydrolase (GH) family 13, specifically hydrolyzes an α-1,6-glucosidic linkage at the non-reducing ends of isomaltooligosaccharides and dextran. SMDG shows the highest sequence similarity to oligo-1,6-glucosidases (O16Gs) among GH family 13 enzymes, but these enzymes are obviously different in terms of substrate chain length specificity. SMDG efficiently hydrolyzes both short-and long-chain substrates, while O16G acts on only short-chain substrates. We focused on this difference in substrate specificity between SMDG and O16G, and elucidated the structure-function relationship of substrate chain length specificity in SMDG. Crystal structure analysis revealed that SMDG consists of three domains, A, B, and C, which are commonly found in other GH family 13 enzymes. The structural comparison between SMDG and O16G from Bacillus cereus indicated that Trp238, spanning subsites +1 and +2, and short βα loop 4, are characteristic of SMDG, and these structural elements are predicted to be important for high activity toward long-chain substrates. The substrate size preference of SMDG was kinetically analyzed using two mutants: (i) Trp238 was replaced by a smaller amino acid, alanine, asparagine or proline; and (ii) short βα loop 4 was exchanged with the corresponding loop of O16G. Mutant enzymes showed lower preference for long-chain substrates than wild-type enzyme, indicating that these structural elements are essential for the high activity toward long-chain substrates, as implied by structural analysis.  相似文献   

2.
本研究用鸟枪法构建了枯草芽孢杆菌(Bacillus subtilis)HB002的基因组文库,经平板法筛选得到了六株能水解合成底物对-硝基苯-α-D-葡萄糖吡喃糖苷的阳性克隆,经鉴定均含克隆了寡聚-1,6-葡萄糖苷酶基因的重组质粒(命名为pHBM001-pHBM006)。选择pHBM003,对其插入片段测序分析,此片段内有一编码561个氨基酸的开放阅读框,该 蛋白质的计算分子量为65.985kD。HB002的寡聚-1,6-葡萄糖苷酶的氨基酸序列与Bacillus sp.和凝结芽孢杆菌(Bacillus coagulans)的寡聚-1,6-葡萄糖苷酶的氨基酸序列一致性分别为81%、67%,相似性分别为89%、79%。从pHBM003中扩增出寡聚-1,6-葡萄糖苷酶基因,克隆到pBV220上,转化大肠杆菌(Escherichia coli)DH5α,得到三个能水解对-硝基苯-α-D-葡萄糖吡喃糖苷的阳性克隆HBM003-1~HBM003-3,将此三个菌株热诱导表达,SDS-PAGE电泳可检测到特异表达的蛋白质,其中HBM003-1、HBM003-2表达的蛋白约66kD,为完整的寡聚-1,6-葡萄糖苷酶,而HBM003-3表达的蛋白质偏小;表达的蛋白质均有寡聚-1,6-葡萄糖苷酶活性。  相似文献   

3.
Okuyama M  Kaneko A  Mori H  Chiba S  Kimura A 《FEBS letters》2006,580(11):2707-2711
Escherichia coli YicI, a member of glycoside hydrolase family (GH) 31, is an alpha-xylosidase, although its amino-acid sequence displays approximately 30% identity with alpha-glucosidases. By comparing the amino-acid sequence of GH 31 enzymes and through structural comparison of the (beta/alpha)(8) barrels of GH 27 and GH 31 enzymes, the amino acids Phe277, Cys307, Phe308, Trp345, Lys414, and beta-->alpha loop 1 of (beta/alpha)(8) barrel of YicI have been identified as elements that might be important for YicI substrate specificity. In attempt to convert YicI into an alpha-glucosidase these elements have been targeted by site-directed mutagenesis. Two mutated YicI, short loop1-enzyme and C307I/F308D, showed higher alpha-glucosidase activity than wild-type YicI. C307I/F308D, which lost alpha-xylosidase activity, was converted into alpha-glucosidase.  相似文献   

4.
An α-glucosidase (HaG) with the following unique properties was isolated from Halomonas sp. strain H11: (i) high transglucosylation activity, (ii) activation by monovalent cations, and (iii) very narrow substrate specificity. The molecular mass of the purified HaG was estimated to be 58 kDa by sodium dodecyl sulfate-polyacrylamide gel electrophoresis (SDS-PAGE). HaG showed high hydrolytic activities toward maltose, sucrose, and p-nitrophenyl α-D-glucoside (pNPG) but to almost no other disaccharides or malto-oligosaccharides higher than trisaccharides. HaG showed optimum activity to maltose at 30°C and pH 6.5. Monovalent cations such as K(+), Rb(+), Cs(+), and NH(4)(+) increased the enzymatic activity to 2- to 9-fold of the original activity. These ions shifted the activity-pH profile to the alkaline side. The optimum temperature rose to 40°C in the presence of 10 mM NH(4)(+), although temperature stability was not affected. The apparent K(m) and k(cat) values for maltose and pNPG were significantly improved by monovalent cations. Surprisingly, k(cat)/K(m) for pNPG increased 372- to 969-fold in their presence. HaG used some alcohols as acceptor substrates in transglucosylation and was useful for efficient synthesis of α-d-glucosylglycerol. The efficiency of the production level was superior to that of the previously reported enzyme Aspergillus niger α-glucosidase in terms of small amounts of by-products. Sequence analysis of HaG revealed that it was classified in glycoside hydrolase family 13. Its amino acid sequence showed high identities, 60%, 58%, 57%, and 56%, to Xanthomonas campestris WU-9701 α-glucosidase, Xanthomonas campestris pv. raphani 756C oligo-1,6-glucosidase, Pseudomonas stutzeri DSM 4166 oligo-1,6-glucosidase, and Agrobacterium tumefaciens F2 α-glucosidase, respectively.  相似文献   

5.
Three active site residues (Asp199, Glu255, Asp329) and two substrate-binding site residues (His103, His328) of oligo-1,6-glucosidase (EC 3.2.1.10) from Bacillus cereus ATCC7064 were identified by site-directed mutagenesis. These residues were deduced from the X-ray crystallographic analysis and the comparison of the primary structure of the oligo-1,6-glucosidase with those of Saccharomyces carlsbergensis alpha-glucosidase, Aspergillus oryzae alpha-amylase and pig pancreatic alpha-amylase which act on alpha-1,4-glucosidic linkages. The distances between these putative residues of B. cereus oligo-1,6-glucosidase calculated from the X-ray analysis data closely resemble those of A. oryzae alpha-amylase and pig pancreatic alpha-amylase. A single mutation of Asp199-->Asn, Glu255-->Gln, or Asp329-->Asn resulted in drastic reduction in activity, confirming that three residues are crucial for the reaction process of alpha-1,6-glucosidic bond cleavage. Thus, it is identified that the basic mechanism of oligo-1,6-glucosidase for the hydrolysis of alpha-1,6-glucosidic linkage is essentially the same as those of other amylolytic enzymes belonging to Family 13 (alpha-amylase family). On the other hand, mutations of histidine residues His103 and His328 resulted in pronounced dissimilarity in catalytic function. The mutation His328-->Asn caused the essential loss in activity, while the mutation His103-->Asn yielded a mutant enzyme that retained 59% of the k0/Km of that for the wild-type enzyme. Since mutants of other alpha-amylases acting on alpha-1,4-glucosidic bond linkage lost most of their activity by the site-directed mutagenesis at their equivalent residues to His103 and His328, the retaining of activity by His103-->Asn mutation in B. cereus oligo-1,6-glucosidase revealed the distinguished role of His103 for the hydrolysis of alpha-1,6-glucosidic bond linkage.  相似文献   

6.
The gene coding for Bacillus cereus ATCC7064 (mesophile) oligo-1,6-glucosidase was cloned within a 2.8-kb SalI-EcoRI fragment of DNA, using the plasmid pUC19 as a vector and Escherichia coli C600 as a host. E. coli C600 bearing the hybrid plasmid pBCE4 accumulated oligo-1,6-glucosidase in the cytoplasm. The cloned enzyme coincided absolutely with B. cereus oligo-1,6-glucosidase in its Mr (65,000), in its electrophoretic behavior on a polyacrylamide gel with or without sodium dodecyl sulfate, in its isoelectric point (4.5), in the temperature dependence of its stability and activity, and in its antigenic determinants. The nucleotide sequence of B. cereus oligo-1,6-glucosidase gene and its flanking regions was determined with both complementary strands of DNA (each 2838 nucleotides). The gene consisted of an open reading frame of 1674 bp commencing with a ATG start codon and followed by a TAA stop codon. The amino acid sequence deduced from the nucleotide sequence predicted a protein of 558 amino acid residues with a Mr of 66,010. The amino acid composition and Mr were comparable with those of B. cereus oligo-1,6-glucosidase. The predicted N-terminal sequence of 10 amino acid residues agreed completely with that of the cloned ligo-1,6-glucosidase. The deduced amino acid sequence of B. cereus oligo-1,6-glucosidase was 72% and 42% similar to those from Bacillus thermoglucosidasius KP1006 (DSM2542, obligate thermophile) oligo-1,6-glucosidase and from Saccharomyces carlsbergensis CB11 alpha-glucosidase, respectively. Predictions of protein secondary structures along with amino acid sequence alignments demonstrated that B. cereus oligo-1,6-glucosidase may take the similar (alpha/beta)8-barrel super-secondary structure, a barrel of eight parallel beta-strands surrounded by eight alpha-helices, in its N-terminal active site domain as S. carlsbergensis alpha-glucosidase and Aspergillus oryzae alpha-amylase.  相似文献   

7.
To identify the critical sites for protein thermostabilization by proline substitution, the gene for oligo-1,6- glucosidase from a thermophilic Bacillus coagulans strain, ATCC 7050, was cloned as a 2.4-kb DNA fragment and sequenced. In spite of a big difference in their thermostabilities, B. coagulans oligo-1,6-glucosidase had a large number of points in its primary structure identical to respective points in the same enzymes from a mesophilic Bacillus cereus strain, ATCC 7064 (57%), and an obligately thermophilic Bacillus thermoglucosidasius strain, KP1006 (59%). The number of prolines (19 for B. cereus oligo-1,6-glucosidase, 24 for B. coagulans enzyme, and 32 for B. thermoglucosidasius enzyme) was observed to increase with the rise in thermostabilities of the oligo-1,6-glucosidases. Classification of proline residues in light of the amino acid sequence alignment and the protein structure revealed by X-ray crystallographic analysis also supported this tendency. Judging from proline residues occurring in B. coagulans oligo-1,6-glucosidase and the structural requirement for proline substitution (second site of the beta turn and first turn of the alpha helix) (K. Watanabe, T. Masuda, H. Ohashi, H. Mihara, and Y. Suzuki, Eur. J. Biochem. 226:277-283, 1994), the critical sites for thermostabilization were found to be Lys-121, Glu-290, Lys-457, and Glu-487 in B. cereus oligo-1,6-glucosidase. With regard to protein evolution, the oligo-1,6-glucosidases very likely follow the neutral theory. The adaptive mutations of the oligo-1,6-glucosidases that appear to increase thermostability are consistent with the substitution of proline residues for neutrally occurring residues. It is concluded that proline substitution is an important factor for the selection of thermostability in oligo-1,6-glucosidases.  相似文献   

8.
Dextran glucosidases show high sequence identity (50%) to Bacillus sp. SAM1606 alpha-glucosidase, which is more specific for short-chain substrates. Sequence comparison of these enzymes as well as molecular modeling studies predicted that the extension of loop 4 of the (beta/alpha)(8)-barrel fold may be responsible for the narrower specificity of SAM1606 alpha-glucosidase with respect to substrate chain length. Indeed, deletion mutants of SAM1606 alpha-glucosidase that lack this extension showed higher relative activities toward dextran and long-chain isomaltooligosaccharides. Kinetic and thermodynamic analyses of oligosaccharide hydrolysis catalyzed by SAM1606 alpha-glucosidase and its deletion mutants suggested that the loss of such extension(s) in loop 4 should energetically destabilize the Michaelis complexes with long-chain substrates to result in smaller differences between the activation free energies for the enzymatic hydrolyses of isomaltoheptaose and isomaltose than those observed for the wild-type enzyme. This is the reason that dextran glucosidase, whose loop 4 is shorter in length, shows broader substrate chain-length specificity than does SAM1606 alpha-glucosidase.  相似文献   

9.
The gene encoding for an extremely thermostable oligo-1,6-glucosidase from Bacillus thermoglucosidasius KP1006 (DSM2542, obligate thermophile) was sequenced. The amino acid sequence deduced from the nucleotide sequence of the gene (1686 base pairs) corresponded to a protein of 562 amino acid residues with a Mr of 66,502. Its predicted amino acid composition, Mr, and N-terminal sequence of 12 residues were consistent with those determined for B. thermoglucosidasius oligo-1,6-glucosidase. The deduced sequence of the enzyme was 72% homologous to that of a thermolabile oligo-1,6-glucosidase (558 residues) from Bacillus cereus ATCC7064 (mesophile). B. cereus oligo-1,6-glucosidase contained 19 prolines. Eighteen of these were conserved at the equivalent positions of B. thermoglucosidasius oligo-1,6-glucosidase. This enzyme contained 14 extra prolines besides the conservative prolines. The majority of extra prolines was replaced by polar or charged residues (Glu, Thr, or Lys) in B. cereus oligo-1,6-glucosidase. The extra prolines were responsible for the difference in thermostability between these two enzymes. We suggested that 11 of the extra prolines in B. thermoglucosidasius oligo-1,6-glucosidase occur in beta-turns or in coils within the loops binding adjacent secondary structures.  相似文献   

10.
Three active site residues (Asp199, Glu255, Asp329) and two substrate-binding site residues (His103, His328) of oligo-1,6-glucosidase (EC 3.2.1.10) from Bacillus cereus ATCC7064 were identified by site-directed mutagenesis. These residues were deduced from the X-ray crystallographic analysis and the comparison of the primary structure of the oligo-1,6-glucosidase with those of Saccharomyces carlsbergensis α-glucosidase, Aspergillus oryzae α-amylase and pig pancreatic α-amylase which act on α-1,4-glucosidic linkages. The distances between these putative residues of B. cereus oligo-1,6-glucosidase calculated from the X-ray analysis data closely resemble those of A. oryzae α-amylase and pig pancreatic α-amylase. A single mutation of Asp199→Asn, Glu255→Gln, or Asp329→Asn resulted in drastic reduction in activity, confirming that three residues are crucial for the reaction process of α-1,6-glucosidic bond cleavage. Thus, it is identified that the basic mechanism of oligo-1,6-glucosidase for the hydrolysis of α-1,6-glucosidic linkage is essentially the same as those of other amylolytic enzymes belonging to Family 13 (α-amylase family). On the other hand, mutations of histidine residues His103 and His328 resulted in pronounced dissimilarity in catalytic function. The mutation His328→Asn caused the essential loss in activity, while the mutation His103→Asn yielded a mutant enzyme that retained 59% of the κ0/Km of that for the wild-type enzyme. Since mutants of other α-amylases acting on α-1,4-glucosidic bond linkage lost most of their activity by the site-directed mutagenesis at their equivalent residues to His103 and His328, the retaining of activity by Hisl03→Asn mutation in B. cereus oligo-1,6-glucosidase revealed the distinguished role of His103 for the hydrolysis of α-1,6-glucosidic bond linkage.  相似文献   

11.
Bacillus stearothermophilus alpha-1,4-glucosidase (BS) is highly specific for alpha-1,4-glucosidic bonds of maltose, maltooligosaccharides and alpha-glucans. Bacillus thermoglucosdasius oligo-1,6-glucosidase (BT) can specifically hydrolyse alpha-1,6 bonds of isomaltose, isomaltooligosaccharides and alpha-limit dextrin. The two enzymes have high homology in primary structure and belong to glycoside hydrolase family 13, which contain four conservative regions (I, II, III and IV). The two enzymes are suggested to be very close in structure, even though there are strict differences in their substrate specificities. Molecular determinants of substrate recognition in these two enzymes were analysed by site-directed mutagenesis. Twenty BT-based mutants and three BS-based mutants were constructed and characterized. Double substitutions in BT of Val200 -->Ala in region II and Pro258 -->Asn in region III caused an appearance of maltase activity compared with BS, and a large reduction of isomaltase activity. The values of k(0)/K(m) (s(-1). mM(-1)) of the BT-mutant for maltose and isomaltose were 69.0 and 15.4, respectively. We conclude that the Val/Ala200 and Pro/Asn258 residues in the alpha-glucosidases may be largely responsible for substrate recognition, although the regions I and IV also exert a slight influence. Additionally, BT V200A and V200A/P258N possessed high hydrolase activity towards sucrose.  相似文献   

12.
Met53 in barley alpha-amylase 1 (AMY1) is situated at the high-affinity subsite -2. While Met53 is unique to plant alpha-amylases, the adjacent Tyr52 stacks onto substrate at subsite -1 and is essentially invariant in glycoside hydrolase family 13. These residues belong to a short sequence motif in beta-->alpha loop 2 of the catalytic (beta/alpha)8-barrel and site-directed mutagenesis was used to introduce a representative variety of structural changes, Met53Glu/Ala/Ser/Gly/Asp/Tyr/Trp, to investigate the role of Met53. Compared to wild-type, Met53Glu/Asp AMY1 displayed 117/90% activity towards insoluble Blue Starch, and Met53Ala/Ser/Gly 76/58/38%, but Met53Tyr/Trp only 0.9/0.1%, even though both Asp and Trp occur frequently at this position in family 13. Towards amylose DP17 (degree of polymerization = 17) and 2-chloro-4-nitrophenyl beta-d-maltoheptaoside the activity (kcat/Km) of all mutants was reduced to 5.5-0.01 and 1.7-0.02% of wild-type, respectively. Km increased up to 20-fold for these soluble substrates and the attack on glucosidic linkages in 4-nitrophenyl alpha-d-maltohexaoside (PNPG6) and PNPG5 was determined by action pattern analysis to shift to be closer to the nonreducing end. This indicated that side chain replacement at subsite -2 weakened substrate glycon moiety contacts. Thus whereas all mutants produced mainly PNPG2 from PNPG6 and similar amounts of PNPG2 and PNPG3 accounting for 85% of the products from PNPG5, wild-type released 4-nitrophenol from PNPG6 and PNPG and PNPG2 in equal amounts from PNPG5. Met53Trp affected the action pattern on PNPG7, which was highly unusual for AMY1 subsite mutants. It was also the sole mutant to catalyze substantial transglycosylation - promoted probably by slow substrate hydrolysis - to produce up to maltoundecaose from PNPG6.  相似文献   

13.
Cultured skin fibroblasts from a patient with I-Cell disease (mucolipidosis II) were assayed for a number of lysosomal enzymes using both natural and synthetic substrates. The cells from this patient were found to have very low activity for galactosylceramide β-galactosidase, lactosylceramide β-galactosidases (using two assay methods that measure different enzymes), GM1 ganglioside β-galactosidase and sphingomyelinase. Glucosylceramide β-glucosidase activity was found to be normal. Acid hydrolase activities toward many synthetic substrate were measured and all except β-glucosidase and acid phosphatase were found to be extremely low (as has been reported by others). Acid phosphatase and β-glucosidase were in the low normal range. These studies expand on previously published reports on I-Cell disease that only present data from synthetic substrates, and also report the fibroblast culture deficiencies of galactosyl-ceramide β-galactosidase (the Krabbe disease enzyme) and sphingomyelinase (the Niemann-Pick disease enzyme) activities for the first time. Those two enzymes do not have a readily available synthetic analog to assay. Acid β-galactosidase activity measured with both the 4-methylumbelliferyl derivative and GM1 ganglioside was partially deficient in leukocytes prepared from this patient. New methods for measuring 4-methylumbelliferyl-β-D-glucoside and glucosylceramide β-glucosidase activities are also presented.  相似文献   

14.
Debranching enzyme was purified from Saccharomyces cerevisiae by DEAE-cellulose, omega-aminobutyl agarose and hydroxyapatite column chromatography. The activity of the eluent was monitored by the iodine-staining method which detects both the direct and indirect debranching enzymes. The elution profiles at every step showed a single peak with no shoulder. The crude and the purified enzyme preparations gave a single activity band with the same mobility on PAGE. The crude product produced 80% glucose compared to reducing sugar from glycogen-phosphorylase-limited dextrin while the partially purified and purified preparations produced 100% glucose. The activity of the purified enzyme was characterized and compared with that of the rabbit muscle enzyme by using various branched cyclodextrins as substrates. Both enzymes hydrolyzed 6-O-alpha-D-glucosyl cyclodextrins to glucose and cyclodextrins, but did not act on 6-O-alpha-maltosyl cyclomaltoheptaose. The yeast enzyme gave rise to glucose as a sole reducing sugar from 6-O-alpha-maltotriosyl cyclomaltoheptaose and 6-O-alpha-maltotetraosyl cyclomaltoheptaose, indicating that maltosyl and maltotriosyl transfers, respectively, had occurred, prior to the action of amylo-1,6-glucosidase. 6-O-alpha-D-Glucosyl cyclomaltoheptaose and 6-O-alpha-D-glucosyl cyclomalto-octaose, respectively, were better substrates than glycogen-phosphorylase-limited dextrin for the yeast and muscle enzymes. The yeast enzyme released glucose at a similar rate from 6-O-alpha-maltotriosyl cyclomaltoheptaose as from 6-O-alpha-maltotetraosyl cyclomaltoheptaose, but considerably lower rates than that from limit dextrin. The yeast debranching enzyme appears to be exclusively oligo-1,4----1,4-glucantransferase-amylo-1,6-glucosidase and does not have isoamylase.  相似文献   

15.
Glycogen debranching enzyme (GDE) in mammals and yeast exhibits α-1,4-transferase and α-1,6-glucosidase activities within a single polypeptide chain and facilitates the breakdown of glycogen by a bi-functional mechanism. Each enzymatic activity of GDE is suggested to be associated with distinct domains; α-1,4-glycosyltransferase activity with the N-terminal domain and α-1,6-glucosidase activity with the C-terminal domain. Here, we present the biochemical features of the GDE from Saccharomyces cerevisiae using the substrate glucose(n)-β-cyclodextrin (Gn-β-CD). The bacterially expressed and purified GDE N-terminal domain (aa 1–644) showed α-1,4-transferase activity on maltotetraose (G4) and G4-β-CD, yielding various lengths of (G)n. Surprisingly, the N-terminal domain also exhibited α-1,6-glucosidase activity against G1-β-CD and G4-β-CD, producing G1 and β-CD. Mutational analysis showed that residues D535 and E564 in the N-terminal domain are essential for the transferase activity but not for the glucosidase activity. These results indicate that the N-terminal domain (1–644) alone has both α-1,4-transferase and the α-1,6-glucosidase activities and suggest that the bi-functional activity in the N-domain may occur via one active site, as observed in some archaeal debranching enzymes.  相似文献   

16.
Alphan alpha-amylase (TVA II) from Thermoactinomyces vulgaris R-47 efficiently hydrolyzes alpha-1,4-glucosidic linkages of pullulan to produce panose in addition to hydrolyzing starch. TVA II also hydrolyzes alpha-1,4-glucosidic linkages of cyclodextrins and alpha-1,6-glucosidic linkages of isopanose. To clarify the basis for this wide substrate specificity of TVA II, we soaked 4(3)-alpha-panosylpanose (4(3)-P2) (a pullulan hydrolysate composed of two panosyl units) into crystals of D325N inactive mutated TVA II. We then determined the crystal structure of TVA II complexed with 4(2)-alpha-panosylpanose (4(2)-P2), which was produced by transglycosylation from 4(3)-P2, at 2.2-A resolution. The shape of the active cleft of TVA II is unique among those of alpha-amylase family enzymes due to a loop (residues 193-218) that is located at the end of the cleft around the nonreducing region and forms a 'dam'-like bank. Because this loop is short in TVA II, the active cleft is wide and shallow around the nonreducing region. It is assumed that this short loop is one of the reasons for the wide substrate specificity of TVA II. While Trp356 is involved in the binding of Glc +2 of the substrate, it appears that Tyr374 in proximity to Trp356 plays two roles: one is fixing the orientation of Trp356 in the substrate-liganded state and the other is supplying the water that is necessary for substrate hydrolysis.  相似文献   

17.

Background

The Theta class glutathione transferase GST T1-1 is a ubiquitously occurring detoxication enzyme. The rat and mouse enzymes have high catalytic activities with numerous electrophilic compounds, but the homologous human GST T1-1 has comparatively low activity with the same substrates. A major structural determinant of substrate recognition is the H-site, which binds the electrophile in proximity to the nucleophilic sulfur of the second substrate glutathione. The H-site is formed by several segments of amino acid residues located in separate regions of the primary structure. The C-terminal helix of the protein serves as a lid over the active site, and contributes several residues to the H-site.

Methods

Site-directed mutagenesis of the H-site in GST T1-1 was used to create the mouse Arg234Trp for comparison with the human Trp234Arg mutant and the wild-type rat, mouse, and human enzymes. The kinetic properties were investigated with an array of alternative electrophilic substrates to establish substrate selectivity profiles for the different GST T1-1 variants.

Results

The characteristic activity profile of the rat and mouse enzymes is dependent on Arg in position 234, whereas the human enzyme features Trp. Reciprocal mutations of residue 234 between the rodent and human enzymes transform the substrate-selectivity profiles from one to the other.

Conclusions

H-site residue 234 has a key role in governing the activity and substrate selectivity profile of GST T1-1.

General significance

The functional divergence between human and rodent Theta class GST demonstrates that a single point mutation can enable or suppress enzyme activities with different substrates.  相似文献   

18.
Transglucosylation activities of spinach α-glucosidase I and IV, which have different substrate specificity for hydrolyzing activity, were investigated. In a maltose mixture, α-glucosidase I, which has high activity toward not only maltooligosaccharides but also soluble starch and can hydrolyze isomaltose, produced maltotriose, isomaltose, and panose, and α-glucosidase IV, which has high activity toward maltooligosaccharides but faint activity toward soluble starch and isomaltose, produced maltotriose, kojibiose, and 2,4-di-α-D-glucosyl-glucose. Transglucosylation to sucrose by α-glucosidase I and IV resulted in the production of theanderose and erlose, respectively, showing that spinach α-glucosidase I and IV are useful to synthesize the α-1,6-glucosylated and α-1,2- and 1,4-glucosylated products, respectively.  相似文献   

19.
The nonreducing end of the substrate-binding site of human salivary alpha-amylase contains two residues Trp58 and Trp59, which belong to beta2-alpha2 loop of the catalytic (beta/alpha)(8) barrel. While Trp59 stacks onto the substrate, the exact role of Trp58 is unknown. To investigate its role in enzyme activity the residue Trp58 was mutated to Ala, Leu or Tyr. Kinetic analysis of the wild-type and mutant enzymes was carried out with starch and oligosaccharides as substrates. All three mutants exhibited a reduction in specific activity (150-180-fold lower than the wild type) with starch as substrate. With oligosaccharides as substrates, a reduction in k(cat), an increase in K(m) and distinct differences in the cleavage pattern were observed for the mutants W58A and W58L compared with the wild type. Glucose was the smallest product generated by these two mutants in the hydrolysis oligosaccharides; in contrast, wild-type enzyme generated maltose as the smallest product. The production of glucose by W58L was confirmed from both reducing and nonreducing ends of CNP-labeled oligosaccharide substrates. The mutant W58L exhibited lower binding affinity at subsites -2, -3 and +2 and showed an increase in transglycosylation activity compared with the wild type. The lowered affinity at subsites -2 and -3 due to the mutation was also inferred from the electron density at these subsites in the structure of W58A in complex with acarbose-derived pseudooligosaccharide. Collectively, these results suggest that the residue Trp58 plays a critical role in substrate binding and hydrolytic activity of human salivary alpha-amylase.  相似文献   

20.
The six forms of the 17alpha-hydroxy steroid dehydrogenase purified from rabbit liver cytosol have very similar physical properties. The molecular weights of all the enzymes were within 3% of the average mol.wt of 39 600. Only one of the six enzymes showed a significant difference in amino acid composition. All but one form of the 17alpha-hydroxy steroid dehydrogenases exhibited greater activities towards the androgen, epitestosterone, than towards oestrogen substrates. With oestrogen substrates one enzyme displayed a high specificity towards the substrate oestradiol-17alpha 3-glucuronide. This high activity was lost if the glucuronic acid moiety was removed or replaced by glucose or galacturonic acid. The other enzyme forms had approximately equal activity toward oestradiol-17alpha and its glucuronide or glucoside derivative. However, substitution of galacturonic acid at C-3 of oestradiol-17alpha substantially decreased the activity of all but one enzyme form.  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