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 515 毫秒
1.
A beta‐galactosidase from the digestive juice of the palm weevil Rhynchophorus palmarum L. larvae was purified by chromatography on ion exchange, gel filtration, and hydrophobic interaction columns. The preparation was shown to be homogeneous on polyacrylamide gel. Beta‐galactosidase was a monomeric protein with a molecular weight of 62 kDa based on its mobility in sodium dodecyl sulfate–polyacrylamide gel electrophoresis and 60 kDa based on gel filtration. Maximal enzyme activity occurred at 55°C and pH 5.0. The purified beta‐galactosidase was stable at 37°C and its pH stability was in the range of 4.6–6.0. Beta‐galactosidase was highly specific for the beta‐d ‐galactosyl residue and beta‐(1‐4) linkage. The catalytic efficiency (Vmax/Km) values for p‐nitrophenyl‐beta‐d ‐galactopyranoside, beta‐d ‐galactosyl(1‐4)‐d ‐glucose (lactose), beta‐d ‐galactosyl(1‐4)‐d ‐galactose and beta‐d ‐galactosyl(1‐4)‐beta‐d ‐galactosyl(1‐4)‐d ‐glucose were, respectively, 72.95, 10.97, 20.74 and 12.73. 5,5‐Dithio‐bis(2‐nitrobenzoate) and sodium dodecyl sulfate inhibited completely the beta‐galactosidase activity. The enzyme was capable of catalyzing transgalactosylation reactions. The yield of galactosylation of 2‐phenylethanol (43%), catalyzed by the beta‐galactosidase in the presence of lactose as galactosyl donor, is higher than those reported previously with conventional sources of beta‐galactosidases. In addition, the optimum pH is different for the hydrolysis (pH 5.0) and transgalactosylation reactions (pH 6.0).  相似文献   

2.
The 50‐residue snake venom protein L ‐omwaprin and its enantiomer D ‐omwaprin were prepared by total chemical synthesis. Radial diffusion assays were performed against Bacillus megaterium and Bacillus anthracis; both L ‐ and D ‐omwaprin showed antibacterial activity against B. megaterium. The native protein enantiomer, made of L ‐amino acids, failed to crystallize readily. However, when a racemic mixture containing equal amounts of L ‐ and D ‐omwaprin was used, diffraction quality crystals were obtained. The racemic protein sample crystallized in the centrosymmetric space group P21/c and its structure was determined at atomic resolution (1.33 Å) by a combination of Patterson and direct methods based on the strong scattering from the sulfur atoms in the eight cysteine residues per protein. Racemic crystallography once again proved to be a valuable method for obtaining crystals of recalcitrant proteins and for determining high‐resolution X‐ray structures by direct methods.  相似文献   

3.
The l ‐ascorbate (AsA) content and the expression of six l ‐galactose pathway‐related genes were analyzed in peach flesh during fruit development. Fluctuation of AsA during peach fruit development was divided into four phases based on the overall total AsA (T‐AsA) content per fruit: AsA I, 0–36 days after full bloom (DAFB); AsA II, 37–65 DAFB; AsA III, 66–92 DAFB and AsA IV, 93–112 DAFB. Phase AsA III was a lag phase for AsA accumulation, but did not coincide with the lag phase for fruit development. The T‐AsA concentration was highest at the early stage until 21 DAFB [2–3μmol per gram of fresh weight (g?1 FW)], and decreased to 1/4 and 1/15 of this value at 50 and 92 DAFB, respectively. T‐AsA then remained at 0.15–0.20μmol g?1 FW until harvest at 112 DAFB. More than 90% of the T‐AsA was in the reduced form until 21 DAFB. The proportion of reduced form of AsA then decreased concomitantly with the decrease in AsA concentration. To determine the main pathway of AsA biosynthesis and the AsA biosynthetic capacity of peach flesh, several precursors were incubated with immature whole fruit (59 DAFB). The AsA concentration increased markedly with l ‐galactono‐1,4‐lactone or l ‐galactose (Gal), but d ‐galacturonate and l ‐gulono‐1,4‐lactone failed to increase AsA, indicating dominance of the Gal pathway and potent AsA biosynthetic capabilities in immature peach flesh. The expression of genes involved in the last six steps of the Gal pathway was measured during fruit development. The genes studied included GDP‐d ‐mannose pyrophosphorylase (GMPH), GDP‐ d ‐mannose‐3′,5′‐epimerase (GME), GDP‐ l ‐galactose guanylyltransferase (GGGT), l ‐galactose‐1‐phosphate phosphatase (GPP), l ‐galactose‐1‐dehydrogenase (GDH) and l ‐galactono‐1,4‐lactone dehydrogenase (GLDH). GMPH, GME and GGGT had similar expression patterns that peaked at 43 DAFB. GPP, GDH and GLDH also had similar expression patterns that peaked twice at 21 and 91 DAFB, although the expression of GDH was quite low. High level of T‐AsA concentration was roughly correlated with the level of gene expression in the early period of fruit development (AsA I), whereas no such relationships were apparent in the other periods (e.g. AsA III and IV). On the basis of these findings, we discuss the regulation of AsA biosynthesis in peach fruit.  相似文献   

4.
Three pathogenic species of the genus Yersinia assemble adhesive fimbriae via the FGL‐chaperone/usher pathway. Closely related Y. pestis and Y. pseudotuberculosis elaborate the pH6 antigen (Psa), which mediates bacterial attachment to alveolar cells of the lung. Y. enterocolitica, instead, assembles the homologous fimbriae Myf of unknown function. Here, we discovered that Myf, like Psa, specifically recognizes β1‐3– or β1‐4–linked galactose in glycosphingolipids, but completely lacks affinity for phosphatidylcholine, the main receptor for Psa in alveolar cells. The crystal structure of a subunit of Psa (PsaA) complexed with choline together with mutagenesis experiments revealed that PsaA has four phosphatidylcholine binding pockets that enable super‐high‐avidity binding of Psa‐fibres to cell membranes. The pockets are arranged as six tyrosine residues, which are all missing in the MyfA subunit of Myf. Conversely, the crystal structure of the MyfA‐galactose complex revealed that the galactose‐binding site is more extended in MyfA, enabling tighter binding to lactosyl moieties. Our results suggest that during evolution, Psa has acquired a tyrosine‐rich surface that enables it to bind to phosphatidylcholine and mediate adhesion of Y. pestis/pseudotuberculosis to alveolar cells, whereas Myf has specialized as a carbohydrate‐binding adhesin, facilitating the attachment of Y. enterocolitica to intestinal cells.  相似文献   

5.
Glyco‐design of proteins is a powerful tool in fundamental studies of structure–function relationship and in obtaining profiles optimized for efficacy of therapeutic glycoproteins. Plants, particularly Nicotiana benthamiana, are attractive hosts to produce recombinant glycoproteins, and recent advances in glyco‐engineering facilitate customized N‐glycosylation of plant‐derived glycoproteins. However, with exception of monoclonal antibodies, homogenous human‐like β1,4‐galactosylation is very hard to achieve in recombinant glycoproteins. Despite significant efforts to optimize the expression of β1,4‐galactosyltransferase, many plant‐derived glycoproteins still exhibit incomplete processed N‐glycans with heterogeneous terminal galactosylation. The most obvious suspects to be involved in trimming terminal galactose residues are β‐galactosidases (BGALs) from the glycosyl hydrolase family GH35. To elucidate the so far uncharacterized mechanisms leading to the trimming of terminal galactose residues from glycans of secreted proteins, we studied a N. benthamiana BGAL known to be active in the apoplast (NbBGAL1). Here, we determined the NbBGAL1 subcellular localization, substrate specificity and in planta biological activity. We show that NbBGAL1 can remove β1,4‐ and β1,3‐galactose residues on both N‐ and O‐glycans. Transient BGAL1 down‐regulation by RNA interference (RNAi) and BGAL1 depletion by genome editing drastically reduce β‐galactosidase activity in N. benthamiana and increase the amounts of fully galactosylated complex N‐glycans on several plant‐produced glycoproteins. Altogether, our data demonstrate that NbBGAL1 acts on galactosylated complex N‐glycans of plant‐produced glycoproteins.  相似文献   

6.
Pectins are the most complex polysaccharides of the plant cell wall. Based on the number of methylations, acetylations and glycosidic linkages present in their structures, it is estimated that up to 67 transferase activities are involved in pectin biosynthesis. Pectic galactans constitute a major part of pectin in the form of side‐chains of rhamnogalacturonan‐I. In Arabidopsis, galactan synthase 1 (GALS1) catalyzes the addition of galactose units from UDP‐Gal to growing β‐1,4‐galactan chains. However, the mechanisms for obtaining varying degrees of polymerization remain poorly understood. In this study, we show that AtGALS1 is bifunctional, catalyzing both the transfer of galactose from UDP‐α‐d ‐Gal and the transfer of an arabinopyranose from UDP‐β‐l ‐Arap to galactan chains. The two substrates share a similar structure, but UDP‐α‐d ‐Gal is the preferred substrate, with a 10‐fold higher affinity. Transfer of Arap to galactan prevents further addition of galactose residues, resulting in a lower degree of polymerization. We show that this dual activity occurs both in vitro and in vivo. The herein described bifunctionality of AtGALS1 may suggest that plants can produce the incredible structural diversity of polysaccharides without a dedicated glycosyltransferase for each glycosidic linkage.  相似文献   

7.
The relationship between the conformation and biological activity of the peptide allosteric modulator of the interleukin‐1 receptor 101.10 (D ‐Arg‐D ‐Tyr‐D ‐Thr‐D ‐Val‐D ‐Glu‐D ‐Leu‐D ‐Ala‐NH2) has been studied using (R)‐ and (S)‐Bgl residues. Twelve Bgl peptides were synthesized using (R)‐ and (S)‐cyclic sulfamidate reagents derived from L ‐ and D ‐aspartic acid in an optimized Fmoc‐compatible protocol for efficient lactam installment onto the supported peptide resin. Examination of these (R)‐ and (S)‐Bgl 101.10 analogs for their potential to inhibit IL‐1β‐induced thymocyte cell proliferation using a novel fluorescence assay revealed that certain analogs exhibited retained and improved potency relative to the parent peptide 101.10. In light of previous reports that Bgl residues may stabilize type II′β‐turn‐like conformations in peptides, CD spectroscopy was performed on selected compounds to identify secondary structure necessary for peptide biological activity. Results indicate that the presence of a fold about the central residues of the parent peptide may be important for activity. Copyright © 2011 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

8.
The globiferous pedicellariae of the venomous sea urchin Toxopneustes pileolus contains several biologically active proteins. We have cloned the cDNA of one of the toxin components, SUL‐I, which is a rhamnose‐binding lectin (RBL) that acts as a mitogen through binding to carbohydrate chains on target cells. Recombinant SUL‐I (rSUL‐I) was produced in Escherichia coli cells, and its carbohydrate‐binding specificity was examined with the glycoconjugate microarray analysis, which suggested that potential target carbohydrate structures are galactose‐terminated N‐glycans. rSUL‐I exhibited mitogenic activity for murine splenocyte cells and toxicity against Vero cells. The three‐dimensional structure of the rSUL‐I/l ‐rhamnose complex was determined by X‐ray crystallographic analysis at a 1.8 Å resolution. The overall structure of rSUL‐I is composed of three distinctive domains with a folding structure similar to those of CSL3, a RBL from chum salmon (Oncorhynchus keta) eggs. The bound l ‐rhamnose molecules are mainly recognized by rSUL‐I through hydrogen bonds between its 2‐, 3‐, and 4‐hydroxy groups and Asp, Asn, and Glu residues in the binding sites, while Tyr and Ser residues participate in the recognition mechanism. It was also inferred that SUL‐I may form a dimer in solution based on the molecular size estimated via dynamic light scattering as well as possible contact regions in its crystal structure.  相似文献   

9.
对脑膜炎败血伊丽莎白金菌进行糖苷酶功能基因组分析发现,该菌存在一个GH27家族α-半乳糖苷酶基因。克隆该基因并表达蛋白后,利用人工合成的pNP底物研究其酶学性质,发现该酶具有α连接的半乳糖(α-galactose,α-Gal)底物特异性,酶反应pH为3.0~8.0,反应温度为4~45 ℃。用不同寡糖底物进一步确定,该酶能够酶切直链末端α-1,3、1,4、1,6Gal。在猪红细胞上进行的酶切实验显示,该酶能够高效清除存在于猪红细胞表面的末端Gal α 1-3Gal表位。末端α-半乳糖基化修饰在免疫与感染中发挥着重要的生物学作用,作用于种属差异性表位的细菌α-半乳糖苷酶的发现将为该领域的研究提供一个新的工具,为缓解异种输血中的超急性免疫排斥反应提供一种可能。  相似文献   

10.
Aims: To characterize of a thermostable recombinant α‐l ‐arabinofuranosidase from Caldicellulosiruptor saccharolyticus for the hydrolysis of arabino‐oligosaccharides to l ‐arabinose. Methods and Results: A recombinant α‐l ‐arabinofuranosidase from C. saccharolyticus was purified by heat treatment and Hi‐Trap anion exchange chromatography with a specific activity of 28·2 U mg?1. The native enzyme was a 58‐kDa octamer with a molecular mass of 460 kDa, as measured by gel filtration. The catalytic residues and consensus sequences of the glycoside hydrolase 51 family of α‐l ‐arabinofuranosidases were completely conserved in α‐l ‐arabinofuranosidase from C. saccharolyticus. The maximum enzyme activity was observed at pH 5·5 and 80°C with a half‐life of 49 h at 75°C. Among aryl‐glycoside substrates, the enzyme displayed activity only for p‐nitrophenyl‐α‐l ‐arabinofuranoside [maximum kcat/Km of 220 m(mol l?1)?1 s?1] and p‐nitrophenyl‐α‐l ‐arabinopyranoside. This substrate specificity differs from those of other α‐l ‐arabinofuranosidases. In a 1 mmol l?1 solution of each sugar, arabino‐oligosaccharides with 2–5 monomer units were completely hydrolysed to l ‐arabinose within 13 h in the presence of 30 U ml?1 of enzyme at 75°C. Conclusions: The novel substrate specificity and hydrolytic properties for arabino‐oligosaccharides of α‐l ‐arabinofuranosidase from C. saccharolyticus demonstrate the potential in the commercial production of l ‐arabinose in concert with endoarabinanase and/or xylanase. Significance and Impact of the Study: The findings of this work contribute to the knowledge of hydrolytic properties for arabino‐oligosaccharides performed by thermostable α‐l ‐arabinofuranosidase.  相似文献   

11.
Many β‐galactosidases show large differences in galacto‐oligosaccharide (GOS) production and lactose hydrolysis. In this study, a kinetic model is developed in which the effect of lactose, glucose, galactose, and oligosaccharides on the oNPG converting activity of various β‐galactosidases is quantified. The use of oNPG as a competing substrate to lactose yields more information than can be obtained by examining only the conversion of lactose itself. The reaction rate with lactose or oligosaccharides as substrate relative to that with water as acceptor is much higher for the β‐galactosidase of Bacillus circulans than the β‐galactosidases of Aspergillus oryzae and Kluyveromyces lactis. In addition, the β‐galactosidase of B.circulans has a high reaction rate with galactose as acceptor, in contrast to those of A. oryzae and K. lactis. The latter two are strongly inhibited by galactose. These differences explain why β‐galactosidase of B. circulans gives higher yields in GOS production than other β‐galactosidases. Many of the reaction rate constants for the β‐galactosidase isoforms of B. circulans increase with increasing molecular weight of the isoform. This indicates that the largest isoform β‐gal‐A is most active in GOS production. However, its hydrolysis rate is also much higher than that of the other isoforms, which results in a faster hydrolysis of oligosaccharides as well. © 2013 American Institute of Chemical Engineers Biotechnol. Prog., 30:38–47, 2014  相似文献   

12.
Galactans containing methylated galactan moieties were extracted from Indian agarophytes, namely, Gracilaria millardetii and Gracilaria textorii growing naturally along the west coast of India. The galactans were treated with α‐amylase to remove floridean starch. These were characterized by Fourier Transform‐Infrared (FT IR), 13C‐Nuclear Magnetic Resonance (C NMR), Gas Chromatography‐Mass Spectrometry (GC MS), Inductively Coupled Plasma spectroscopy (ICP) and Gel Permeation Chromatography (GPC) and were found to be composed of d ‐galactose, 6‐O‐methyl‐D‐galactose and 3,6‐anhydro‐L‐galactose. Methylation analyses revealed that both the polysaccharides consisted of 3‐, 2,3‐, 4,6‐linked galactose as well as four‐linked 3,6‐anhydro galactose residues. Both the Gracilaria species produced low gelling (<100 g cm?2) and highly sulfated (2.1% to 4.8%) galactans containing very low heavy metal contents (ICP). These galactans may be of potential utility in food and biological applications.  相似文献   

13.
Wall polysaccharide chemistry varies phylogenetically, suggesting a need for variation in wall enzymes. Although plants possess the genes for numerous putative enzymes acting on wall carbohydrates, the activities of the encoded proteins often remain conjectural. To explore phylogenetic differences in demonstrable enzyme activities, we extracted proteins from 57 rapidly growing plant organs with three extractants, and assayed their ability to act on six oligosaccharides ‘modelling’ selected cell‐wall polysaccharides. Based on reaction products, we successfully distinguished exo‐ and endo‐hydrolases and found high taxonomic variation in all hydrolases screened: β‐d ‐xylosidase, endo‐(1→4)‐β‐d ‐xylanase, β‐d ‐mannosidase, endo‐(1→4)‐β‐d ‐mannanase, α‐d ‐xylosidase, β‐d ‐galactosidase, α‐l ‐arabinosidase and α‐l ‐fucosidase. The results, as GHATAbase, a searchable compendium in Excel format, also provide a compilation for selecting rich sources of enzymes acting on wall carbohydrates. Four of the hydrolases were accompanied, sometimes exceeded, by transglycosylase activities, generating products larger than the substrate. For example, during β‐xylosidase assays on (1→4)‐β‐d ‐xylohexaose (Xyl6), Marchantia, Selaginella and Equisetum extracts gave negligible free xylose but approximately equimolar Xyl5 and Xyl7, indicating trans‐β‐xylosidase activity, also found in onion, cereals, legumes and rape. The yield of Xyl9 often exceeded that of Xyl7–8, indicating that β‐xylanase was accompanied by an endotransglycosylase activity, here called trans‐β‐xylanase, catalysing the reaction 2Xyl6→ Xyl3 + Xyl9. Similar evidence also revealed trans‐α‐xylosidase, trans‐α‐arabinosidase and trans‐α‐arabinanase activities acting on xyloglucan oligosaccharides and (1→5)‐α‐l ‐arabino‐oligosaccharides. In conclusion, diverse plants differ dramatically in extractable enzymes acting on wall carbohydrate, reflecting differences in wall polysaccharide composition. Besides glycosidase and glycanase activities, five new transglycosylase activities were detected. We propose that such activities function in the assembly and re‐structuring of the wall matrix.  相似文献   

14.
The D ‐aldohexose dehydrogenase from the thermoacidophilic archaeon Thermoplasma acidophilum (AldT) is a homotetrameric enzyme that catalyzes the oxidation of several D ‐aldohexoses, especially D ‐mannose. AldT comprises a unique C‐terminal tail motif (residues 247–255) that shuts the active‐site pocket of the neighboring subunit. The functional role of the C‐terminal tail of AldT has been investigated using mutational and crystallographic analyses. A total of four C‐terminal deletion mutants (Δ254, Δ253, Δ252, and Δ249) and two site‐specific mutants (Y86G and P254G) were expressed by Escherichia coli and purified. Enzymatic characterization of these mutants revealed that the C‐terminal tail is a requisite and that the interaction between Tyr86 and Pro254 is critical for enzyme activity. The crystal structure of the Δ249 mutant was also determined. The structure showed that the active‐site loops undergo a significant conformational change, which leads to the structural deformation of the substrate‐binding pocket. Proteins 2009. © 2008 Wiley‐Liss, Inc.  相似文献   

15.
Enzymatic control of the accumulation of verbascose in pea seeds   总被引:4,自引:0,他引:4  
Verbascose, the pentasaccharide of the raffinose family of oligosaccharides, consists of galactose units joined to sucrose. In pea (Pisum sativum) seeds, the content of verbascose is highly variable. In a previous study on a high‐verbascose pea cultivar, the present authors have demonstrated that verbascose is synthesized by a multifunctional stachyose synthase (EC 2.4.1.67), which utilizes raffinose as well as stachyose as a galactosyl acceptor. Herein the results of a study of the cloning and functional expression of stachyose synthase from the low‐verbascose genotype SD1 are reported and it is demonstrated that this line contains a protein with a reduced ability to synthesize verbascose. Analysis of seeds from seven pea lines revealed a positive correlation between verbascose synthase activity and verbascose content. Among these genotypes, only the SD1 line showed low verbascose synthase activity when the data were normalized to stachyose synthase activity. These results suggest that differences in the level of verbascose synthase activity could be caused by mutations in the stachyose synthase gene as well as by variation in the amount of the protein. The lines were also analysed for activity of α‐galactosidase, a catabolic enzyme that could limit the extent of verbascose accumulation. No relationship was found between α‐galactosidase activity and the amount of raffinose family oligosaccharides.  相似文献   

16.
Aim: The present study was aimed at evaluating the effects of the three crucial factors, galactose concentration, inoculum size and moisture content, on α‐galactosidase production by the filamentous actinobacterium Streptomyces griseoloalbus in solid‐state fermentation. Methods and Results: Central Composite design was adopted to derive a statistical model for the optimization of fermentation conditions. Maximum α‐galactosidase yield (117 U g–1 of dry fermented substrate) was obtained when soya bean flour supplemented with 1·5% galactose and with initial moisture content of 40% was inoculated with 1·9 × 106 CFU g?1 initial dry substrate. Conclusions: The model was valid and could result in considerably enhanced enzyme yield. Significance and Impact of the Study: The results indicated a cost effective method for the production of α‐galactosidase using soya bean flour. This is the first report on exploitation of the potential of filamentous bacterium for the production of α‐galactosidase, an enzyme having versatile applications.  相似文献   

17.
Two sulfated fucoidan fractions (Lj3 and Lj5) were extracted from Saccharina japonica and then subjected to acid hydrolysis to obtain Lj3h and Lj5h. Lj3h and Lj5h were characterized using IR, methylation analysis, and mass spectrometry. It was found that Lj3h and Lj5h were homogeneous low molecular weight fucoidans. Specifically, Lj3h was composed of the main chain of 1,3‐linked α‐L‐fucopyranose residues with sulfate at C‐2 and/or C‐4 and three different monosaccharides (galactose, glucose, mannose) branched at C‐2 and/or C‐4 of fucose residue. Lj5h contained backbones of alternating galactopyranose residues and fucopyranose residues attached via a 1→3 linkage (galactofucan) and 1→6 linked galactan. The sulfation pattern was mainly located at C2/C4 fucose or galactose residues and more branches occupied at C‐4 of fucose residue and C‐2, C‐3 or/and C‐6 of galactose residue. In vitro assay indicated that, among the four fucoidans tested, only Lj5 showed potent α‐glucosidase inhibitory activity with IC50 of 153.27±22.89 μg/mL, and the two parent fucoidans, Lj3 and Lj5, showed better antioxidant activity than their derivatives. These findings highlight the structure and bioactivity diversity of Saccharina japonica‐derived fucoidans.  相似文献   

18.
The use of heterogeneous biocatalysis in industrial applications is advantageous and the enzyme stability improvement is a continuous challenge. Therefore, we designed β‐galactosidase heterogeneous biocatalysts by immobilization, involving the support synthesis and enzyme selection (from Bacillus circulans, Kluyveromyces lactis, and Aspergillus oryzae). The underivatized, tailored, macro‐mesoporous silica exhibited high surface area, offered high enzyme immobilization yields and activity. Its chemical activation with glyoxyl groups bound the enzyme covalently, which suppressed lixiviation and conferred higher pH and thermal stability (120‐fold than for the soluble enzyme), without observable reduction of activity/stability due to the presence of silica. The best balance between the immobilization yield (68%), activity (48%), and stability was achieved for Bacillus circulans β‐galactosidase immobilized on glyoxyl‐activated silica, without using stabilizing agents or modifying the enzyme. The enzyme stabilization after immobilization in glyoxyl‐activated silica was similar to that observed in macroporous agarose‐glyoxyl support, with the reported microbiological and mechanical advantages of inorganic supports. The whey lactolysis at pH 6.0 and 25°C by using this catalyst (1 mg ml?1, 290 UI g?1) was still 90%, even after 10 cycles of 10 min, in batch process but it could be also implemented on continuous processes at industrial level with similar results.  相似文献   

19.
The tyrosinase gene from Ralstonia solanacearum (GenBank NP518458) was subjected to random mutagenesis resulting in tyrosinase variants (RVC10 and RV145) with up to 3.2‐fold improvement in kcat, 5.2‐fold lower Km and 16‐fold improvement in catalytic efficiency for D ‐tyrosine. Based on RVC10 and RV145 mutated sequences, single mutation variants were generated with all variants showing increased kcat for D ‐tyrosine compared to the wild type (WT). All single mutation variants based on RV145 had a higher kcat and Km value compared to the RV145 and thus the combination of four mutations in RV145 was antagonistic for turnover, but synergistic for affinity of the enzyme for D ‐tyrosine. Single mutation variant 145_V153A exhibited the highest (6.9‐fold) improvement in kcat and a 2.4‐fold increase in Km compared to the WT. Two single mutation variants, C10_N322S and C10_T183I reduced the Km up to 2.6‐fold for D ‐tyrosine but one variant 145_V153A increased the Km 2.4‐fold compared to the WT. Homology based modeling of R. solanacearum tyrosinase showed that mutation V153A disrupts the van der Waals interactions with an α‐helix providing one of the conserved histidine residues of the active site. The kcat and Km values for L ‐tyrosine decreased for RV145 and RVC10 compared to the WT. RV145 exhibited a 2.1‐fold high catalytic efficiency compared to the WT which is a 7.6‐fold lower improvement compared to D ‐tyrosine. RV145 exhibited a threefold higher monophenolase:diphenolase activity ratio for D ‐tyrosine:D ‐DOPA and a 1.4‐fold higher L ‐tyrosine:L ‐DOPA activity ratio compared to the WT. Biotechnol. Bioeng. 2013; 110: 1849–1857. © 2013 Wiley Periodicals, Inc.  相似文献   

20.
This review provides an overview of the structure, function, and catalytic mechanism of lacZ β‐galactosidase. The protein played a central role in Jacob and Monod's development of the operon model for the regulation of gene expression. Determination of the crystal structure made it possible to understand why deletion of certain residues toward the amino‐terminus not only caused the full enzyme tetramer to dissociate into dimers but also abolished activity. It was also possible to rationalize α‐complementation, in which addition to the inactive dimers of peptides containing the “missing” N‐terminal residues restored catalytic activity. The enzyme is well known to signal its presence by hydrolyzing X‐gal to produce a blue product. That this reaction takes place in crystals of the protein confirms that the X‐ray structure represents an active conformation. Individual tetramers of β‐galactosidase have been measured to catalyze 38,500 ± 900 reactions per minute. Extensive kinetic, biochemical, mutagenic, and crystallographic analyses have made it possible to develop a presumed mechanism of action. Substrate initially binds near the top of the active site but then moves deeper for reaction. The first catalytic step (called galactosylation) is a nucleophilic displacement by Glu537 to form a covalent bond with galactose. This is initiated by proton donation by Glu461. The second displacement (degalactosylation) by water or an acceptor is initiated by proton abstraction by Glu461. Both of these displacements occur via planar oxocarbenium ion‐like transition states. The acceptor reaction with glucose is important for the formation of allolactose, the natural inducer of the lac operon.  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