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.
Ficko-Blean E  Stuart CP  Boraston AB 《Proteins》2011,79(10):2771-2777
CPF_2247 from Clostridium perfringens ATCC 13124 was identified as a putative carbohydrate‐active enzyme by its low sequence identity to endo‐β‐1,4‐glucanases belonging to family 8 of the glycoside hydrolase classification. The X‐ray crystal structure of CPF_2247 determined to 2.0 Å resolution by single‐wavelength anomalous dispersion using seleno‐methionine‐substituted protein revealed an (α/α)6 barrel fold. A large cleft on the surface of the protein contains residues that are structurally conserved with key elements of the catalytic machinery in clan GH‐M glycoside hydrolases. Assessment of CPF_2247 as a carbohydrate‐active enzyme disclosed α‐glucanase activity on amylose, glycogen, and malto‐oligosaccharides. Proteins 2011;. © 2011 Wiley‐Liss, Inc.  相似文献   

2.
Glycoside phosphorylases (GPs) with specificity for β-(1 → 3)-gluco-oligosaccharides are potential candidate biocatalysts for oligosaccharide synthesis. GPs with this linkage specificity are found in two families thus far—glycoside hydrolase family 94 (GH94) and the recently discovered glycoside hydrolase family 149 (GH149). Previously, we reported a crystallographic study of a GH94 laminaribiose phosphorylase with specificity for disaccharides, providing insight into the enzyme's ability to recognize its' sugar substrate/product. In contrast to GH94, characterized GH149 enzymes were shown to have more flexible chain length specificity, with preference for substrate/product with higher degree of polymerization. In order to advance understanding of the specificity of GH149 enzymes, we herein solved X-ray crystallographic structures of GH149 enzyme Pro_7066 in the absence of substrate and in complex with laminarihexaose (G6). The overall domain organization of Pro_7066 is very similar to that of GH94 family enzymes. However, two additional domains flanking its catalytic domain were found only in the GH149 enzyme. Unexpectedly, the G6 complex structure revealed an oligosaccharide surface binding site remote from the catalytic site, which, we suggest, may be associated with substrate targeting. As such, this study reports the first structure of a GH149 phosphorylase enzyme acting on β-(1 → 3)-gluco-oligosaccharides and identifies structural elements that may be involved in defining the specificity of the GH149 enzymes.  相似文献   

3.
Glycogen serves as major energy storage in most living organisms. GlgX, with its gene in the glycogen degradation operon, functions in glycogen catabolism by selectively catalyzing the debranching of polysaccharide outer chains in bacterial glycosynthesis. GlgX hydrolyzes α‐1,6‐glycosidic linkages of phosphorylase‐limit dextrin containing only three or four glucose subunits produced by glycogen phosphorylase. To understand its mechanism and unique substrate specificity toward short branched α‐polyglucans, we determined the structure of GlgX from Escherichia Coli K12 at 2.25 Å resolution. The structure reveals a monomer consisting of three major domains with high structural similarity to the subunit of TreX, the oligomeric bifunctional glycogen debranching enzyme (GDE) from Sulfolobus. In the overlapping substrate binding groove, conserved residues Leu270, Asp271, and Pro208 block the cleft, yielding a shorter narrow GlgX cleft compared to that of TreX. Residues 207–213 form a unique helical conformation that is observed in both GlgX and TreX, possibly distinguishing GDEs from isoamylases and pullulanases. The structural feature observed at the substrate binding groove provides a molecular explanation for the unique substrate specificity of GlgX for G4 phosphorylase‐limit dextrin and the discriminative activity of TreX and GlgX toward substrates of varying lengths. Proteins 2010. © 2010 Wiley‐Liss, Inc.  相似文献   

4.
Recent metagenomic analyses have identified uncultured bacteria that are abundant in the rumen of herbivores and that possess putative biomass-converting enzyme systems. Here we investigate the saccharolytic capabilities of a polysaccharide utilization locus (PUL) that has been reconstructed from an uncultured Bacteroidetes phylotype (SRM-1) that dominates the rumen microbiome of Arctic reindeer. Characterization of the three PUL-encoded outer membrane glycoside hydrolases was performed using chromogenic substrates for initial screening, followed by detailed analyses of products generated from selected substrates, using high-pressure anion-exchange chromatography with electrochemical detection. Two glycoside hydrolase family 5 (GH5) endoglucanases (GH5_g and GH5_h) demonstrated activity against β-glucans, xylans, and xyloglucan, whereas GH5_h and the third enzyme, GH26_i, were active on several mannan substrates. Synergy experiments examining different combinations of the three enzymes demonstrated limited activity enhancement on individual substrates. Binding analysis of a SusE-positioned lipoprotein revealed an affinity toward β-glucans and, to a lesser extent, mannan, but unlike the two SusD-like lipoproteins previously characterized from the same PUL, binding to cellulose was not observed. Overall, these activities and binding specificities correlated well with the glycan content of the reindeer rumen, which was determined using comprehensive microarray polymer profiling and showed an abundance of various hemicellulose glycans. The substrate versatility of this single PUL putatively expands our perceptions regarding PUL machineries, which so far have demonstrated gene organization that suggests one cognate PUL for each substrate type. The presence of a PUL that possesses saccharolytic activity against a mixture of abundantly available polysaccharides supports the dominance of SRM-1 in the Svalbard reindeer rumen microbiome.  相似文献   

5.
GlgE is a recently identified (1→4)-α-d-glucan:phosphate α-d-maltosyltransferase involved in α-glucan biosynthesis in bacteria and is a genetically validated anti-tuberculosis drug target. It is a member of the GH13_3 CAZy subfamily for which no structures were previously known. We have solved the structure of GlgE isoform I from Streptomyces coelicolor and shown that this enzyme has the same catalytic and very similar kinetic properties to GlgE from Mycobacterium tuberculosis. The S. coelicolor enzyme forms a homodimer with each subunit comprising five domains, including a core catalytic α-amylase-type domain A with a (β/α)(8) fold. This domain is elaborated with domain B and two inserts that are specifically configured to define a well conserved donor pocket capable of binding maltose. Domain A, together with domain N from the neighboring subunit, forms a hydrophobic patch that is close to the maltose-binding site and capable of binding cyclodextrins. Cyclodextrins competitively inhibit the binding of maltooligosaccharides to the S. coelicolor enzyme, showing that the hydrophobic patch overlaps with the acceptor binding site. This patch is incompletely conserved in the M. tuberculosis enzyme such that cyclodextrins do not inhibit this enzyme, despite acceptor length specificity being conserved. The crystal structure reveals two further domains, C and S, the latter being a helix bundle not previously reported in GH13 members. The structure provides a framework for understanding how GlgE functions and will help guide the development of inhibitors with therapeutic potential.  相似文献   

6.
J Takrama  N B Madsen 《Biochemistry》1988,27(9):3308-3314
The binding of glucose and a series of oligosaccharides to glycogen debranching enzyme was determined by the ability of the saccharides to decrease the rate of reaction of sulfhydryl groups with 5,5'-dithiobis(2-nitrobenzoate) (DTNB). At pH 7.2, the strength of binding increases with chain length from glucose to maltotriose to maltopentaose but not to maltohexaose, and the free energies for binding of the oligosaccharides suggest subsites of equivalent affinities for the four glucose units following the initial reducing moiety. The rate of reaction of DTNB with enzyme saturated with saccharide is the same for all compounds, suggesting that all the saccharides, including glucose, induce the same conformational state. The site of binding may be that which binds the alpha-1,6-linked side chain of the natural limit dextrin substrate. At pH 8.0, this site exhibits similar characteristics, but an additional site, which may bind the four terminal glucose units of the main chain of the natural substrate, is manifested and exhibits different characteristics, including a very low affinity for glucose itself. The binding of glycogen to the debranching enzyme was monitored by centrifugal separation from the protein and exhibits a much lower dissociation constant than that for the oligomers, suggesting that branched polymers have more than one set of subsites.  相似文献   

7.
Species of the fungal genus Aspergillus are significant human and agricultural pathogens that are often refractory to existing antifungal treatments. Protein farnesyltransferase (FTase), a critical enzyme in eukaryotes, is an attractive potential target for antifungal drug discovery. We report high‐resolution structures of A. fumigatus FTase (AfFTase) in complex with substrates and inhibitors. Comparison of structures with farnesyldiphosphate (FPP) bound in the absence or presence of peptide substrate, corresponding to successive steps in ordered substrate binding, revealed that the second substrate‐binding step is accompanied by motions of a loop in the catalytic site. Re‐examination of other FTase structures showed that this motion is conserved. The substrate‐ and product‐binding clefts in the AfFTase active site are wider than in human FTase (hFTase). Widening is a consequence of small shifts in the α‐helices that comprise the majority of the FTase structure, which in turn arise from sequence variation in the hydrophobic core of the protein. These structural effects are key features that distinguish fungal FTases from hFTase. Their variation results in differences in steady‐state enzyme kinetics and inhibitor interactions and presents opportunities for developing selective anti‐fungal drugs by exploiting size differences in the active sites. We illustrate the latter by comparing the interaction of ED5 and Tipifarnib with hFTase and AfFTase. In AfFTase, the wider groove enables ED5 to bind in the presence of FPP, whereas in hFTase it binds only in the absence of substrate. Tipifarnib binds similarly to both enzymes but makes less extensive contacts in AfFTase with consequently weaker binding.  相似文献   

8.
Starch metabolism in Dunaliella parva Lerche is regulated by the osmolarity of the surrounding solute. Two isozymes showing amylolytic activity were obtained after purification by gel filtration chromatography. The isozymes show disproportionating activity (D‐enzyme) that is specific for malto‐oligosaccharides as substrate. Properties of the D‐enzyme in D. parva are similar to those in higher plants. The activity of the D‐enzyme is also found in various Dunaliella and Chlamydomonas, indicating that the D‐enzyme is also important in the starch metabolism in algae.  相似文献   

9.
Weikl TR  Boehr DD 《Proteins》2012,80(10):2369-2383
Protein function often involves changes between different conformations. Central questions are how these conformational changes are coupled to the binding or catalytic processes during which they occur, and how they affect the catalytic rates of enzymes. An important model system is the enzyme dihydrofolate reductase (DHFR) from Escherichia coli, which exhibits characteristic conformational changes of the active‐site loop during the catalytic step and during unbinding of the product. In this article, we present a general kinetic framework that can be used (1) to identify the ordering of events in the coupling of conformational changes, binding, and catalysis and (2) to determine the rates of the substeps of coupled processes from a combined analysis of nuclear magnetic resonance R2 relaxation dispersion experiments and traditional enzyme kinetics measurements. We apply this framework to E. coli DHFR and find that the conformational change during product unbinding follows a conformational‐selection mechanism, that is, the conformational change occurs predominantly prior to unbinding. The conformational change during the catalytic step, in contrast, is an induced change, that is, the change occurs after the chemical reaction. We propose that the reason for these conformational changes, which are absent in human and other vertebrate DHFRs, is robustness of the catalytic rate against large pH variations and changes to substrate/product concentrations in E. coli. Proteins 2012;. © 2012 Wiley Periodicals, Inc.  相似文献   

10.
Glycoside hydrolase family 77 (GH77) belongs to the alpha-amylase superfamily (Clan H) together with GH13 and GH70. GH77 enzymes are amylomaltases or 4-alpha-glucanotransferases, involved in maltose metabolism in microorganisms and in starch biosynthesis in plants. Here we characterized the amylomaltase from the hyperthermophilic bacterium Thermus thermophilus HB8 (Tt AMase). Site-directed mutagenesis of the active site residues (Asp293, nucleophile; Glu340, general acid/base catalyst; Asp395, transition state stabilizer) shows that GH77 Tt AMase and GH13 enzymes share the same catalytic machinery. Quantification of the enzyme's transglycosylation and hydrolytic activities revealed that Tt AMase is among the most efficient 4-alpha-glucanotransferases in the alpha-amylase superfamily. The active site contains at least seven substrate binding sites, subsites -2 and +3 favoring substrate binding and subsites -3 and +2 not, in contrast to several GH13 enzymes in which subsite +2 contributes to oligosaccharide binding. A model of a maltoheptaose (G7) substrate bound to the enzyme was used to probe the details of the interactions of the substrate with the protein at acceptor subsites +2 and +3 by site-directed mutagenesis. Substitution of the fully conserved Asp249 with a Ser in subsite +2 reduced the activity 23-fold (for G7 as a substrate) to 385-fold (for maltotriose). Similar mutations reduced the activity of alpha-amylases only up to 10-fold. Thus, the characteristics of acceptor subsite +2 represent a main difference between GH13 amylases and GH77 amylomaltases.  相似文献   

11.
BackgroundEnvironmental bacteria express a wide diversity of glycoside hydrolases (GH). Screening and characterization of GH from metagenomic sources provides an insight into biomass degradation strategies of non-cultivated prokaryotes.MethodsIn the present report, we screened a compost metagenome for lignocellulolytic activities and identified six genes encoding enzymes belonging to family GH9 (GH9a-f). Three of these enzymes (GH9b, GH9d and GH9e) were successfully expressed and characterized.ResultsA phylogenetic analysis of the catalytic domain of pro- and eukaryotic GH9 enzymes suggested the existence of two major subgroups. Bacterial GH9s displayed a wide variety of modular architectures and those harboring an N-terminal Ig-like domain, such as GH9b and GH9d, segregated from the remainder. We purified and characterized GH9 endoglucanases from both subgroups and examined their stabilities, substrate specificities and product profiles. GH9e exhibited an original hydrolysis pattern, liberating an elevated proportion of oligosaccharides longer than cellobiose. All of the enzymes exhibited processive behavior and a synergistic action on crystalline cellulose. Synergy was also evidenced between GH9d and a GH48 enzyme identified from the same metagenome.ConclusionsThe characterized GH9 enzymes displayed different modular architectures and distinct substrate and product profiles. The presence of a cellulose binding domain was shown to be necessary for binding and digestion of insoluble cellulosic substrates, but not for processivity.General significanceThe identification of six GH9 enzymes from a compost metagenome and the functional variety of three characterized members highlight the importance of this enzyme family in bacterial biomass deconstruction.  相似文献   

12.
Single molecules of glycogen phosphorylase b exhibit images in the electron microscope which are similar in shape and dimension to those derived from X-ray crystallography. Phosphorylase alpha exhibits tetramers but shows dimers in the presence of glucose. Glycogen debranching enzyme appears as a monomer with an unusual crescent or shrimp-like shape, with occasional isologous aggregation to circular dimers. The longest dimension of the monomer is very similar to that of the phosphorylase dimer, 11.5 nm. Strong binding of the debranching enzyme to glycogen is readily visualized in the electron microscope. It is suggested that the distinctive shape of the debranching enzyme may be related to its catalytic function.  相似文献   

13.
Characterization of Starch-Debranching Enzymes in Pea Embryos   总被引:5,自引:0,他引:5       下载免费PDF全文
Two distinct types of debranching enzymes have been identified in developing pea (Pisum sativum L.) embryos using native gel analysis and tests of substrate preference on purified or partially purified activities. An isoamylase-like activity capable of hydrolyzing amylopectin and glycogen but not pullulan is present throughout development and is largely or entirely confined to the plastid. Activities capable of hydrolyzing pullulan are present both inside and outside of the plastid, and extraplastidial activity increases relative to the plastidial activity during development. Both types of debranching enzyme are also present in germinating embryos. We argue that debranching enzymes are likely to have a role in starch metabolism in the plastid of the developing embryo and in starch degradation during germination.  相似文献   

14.
The polysaccharide utilization locus in Bacteroides plebeius that confers the ability to catabolize porphyran contains a putative GH50 β‐agarase (BACPLE_01683, BpGH50). BpGH50 did not show any clear activity on agarose or on the related algal galactans porphyran and carrageenan. However, the 1.4 Å resolution X‐ray crystal structure of BpGH50 confirmed its possession of the core (α/β)8 barrel fold found in GH50 enzymes as well as the structural conservation of the catalytic residues and some substrate binding residues. Examination of the structure supports assignment of this protein as a β‐galactosidase but suggests that it may utilize a different, possibly hybrid, algal galactan substrate. Proteins 2016; 85:182–187. © 2016 Wiley Periodicals, Inc.  相似文献   

15.
Six glycoside hydrolase (GH) family 13 members, classified under the polyspecific neopullulanase subfamily GH13_20 (also termed cyclomaltodextrinase) were analysed. They originate from thermophilic bacterial strains (Anoxybacillus flavithermus, Laceyella sacchari, and Geobacillus thermoleovorans) or from environmental DNA, collected after in situ enrichments in Icelandic hot springs. The genes were isolated following the CODEHOP consensus primer strategy, utilizing the first two of the four conserved sequence regions in GH13. The typical domain structure of GH13_20, including an N-terminal domain (classified as CBM34), the catalytic module composed of the A-and B-domains, and a C-terminal domain, was found in five of the encoded enzymes (abbreviated Amy1, 89, 92, 98 and 132). These five enzymes degraded cyclomaltodextrins (CDs) and starch, while only three, Amy92 (L. sacchari), Amy98 (A. flavithermus) and Amy132 (environmental DNA), also harboured neopullulanase activity. The L. sacchari enzyme was monomeric, but with CD as the preferred substrate, which is an unusual combination. The sixth enzyme (Amy29 from environmental DNA), was composed of the ABC-domains only. Preferred substrate for Amy29 was pullulan, which was degraded to panose, and the enzyme had no detectable activity on CDs. In addition to its different activity profile and domain composition, Amy29 also displayed a different conservation (LPKF) in the fifth conserved region (MPKL) proposed to identify the subfamily. All enzymes had apparent temperature optima in the range 50–65°C, while thermostability varied, and was highest for Amy29 with a half-life of 480 min at 80°C. Calcium dependent activity or stability was monitored in four enzymes, but could not be detected for Amy29 or 98. Tightly bound calcium can, however, not be ruled out, and putative calcium ligands were conserved in Amy98.  相似文献   

16.
A huge number of glycoside hydrolases are classified into the glycoside hydrolase family (GH family) based on their amino-acid sequence similarity. The glycoside hydrolases acting on α-glucosidic linkage are in GH family 4, 13, 15, 31, 63, 97, and 122. This review deals mainly with findings on GH family 31 and 97 enzymes. Research on two GH family 31 enzymes is described: clarification of the substrate recognition of Escherichia coli α-xylosidase, and glycosynthase derived from Schizosaccharomyces pombe α-glucosidase. GH family 97 is an aberrant GH family, containing inverting and retaining glycoside hydrolases. The inverting enzyme in GH family 97 displays significant similarity to retaining α-glycosidases, including GH family 97 retaining α-glycosidase, but the inverting enzyme has no catalytic nucleophile residue. It appears that a catalytic nucleophile has been eliminated during the molecular evolution in the same way as a man-made nucleophile mutant enzyme, which catalyzes the inverting reaction, as in glycosynthase and chemical rescue.  相似文献   

17.
A gene, treX, encoding a debranching enzyme previously cloned from the trehalose biosynthesis gene cluster of Sulfolobus solfataricus P2 was expressed in Escherichia coli as a His-tagged protein and the biochemical properties were studied. The specific activity of the S. solfataricus debranching enzyme (TreX) was highest at 75°C and pH 5.5. The enzyme exhibited hydrolysing activity toward α-1,6-glycosidic linkages of amylopectin, glycogen, pullulan, and other branched substrates, and glycogen was the preferred substrate. TreX has a high specificity for hydrolysis of maltohexaosyl α-1,6-β-cyclodextrin, indicating the high preference for side chains consisting of 6 glucose residues or more. The enzyme also exhibited 4-α-sulfoxide-glucan transferase activity, catalysing transfer of α-1,4-glucan oligosaccharides from one chain to another. Dimethyl sulfoxide (10%, v/v) increased the hydrolytic activity of TreX. Gel permeation chromatography and sedimentation equilibrium analytical ultracentrifugation revealed that the enzyme exists mostly as a dimer at pH 7.0, and as a mixture of dimers and tetramers at pH 5.5. Interestingly, TreX existed as a tetramer in the presence of DMSO at pH 5.5–6.5. The tetramer showed a 4-fold higher catalytic efficiency than the dimer. The enzyme catalysed not only intermolecular trans-glycosylation of malto-oligosaccharides (disproportionation) to produce linear α-1,4-glucans, but also intramolecular trans-glycosylation of glycogen. The results presented in this study indicated that TreX may be associated with glycogen metabolism by selective cleavage of the outer side chain.  相似文献   

18.
Isoprimeverose-producing enzymes (IPases) release isoprimeverose (α-d-xylopyranosyl-(1?→?6)-d-glucopyranose) from the non-reducing end of xyloglucan oligosaccharides. Aspergillus oryzae IPase (IpeA) is classified as a member of the glycoside hydrolase family 3 (GH3); however, it has unusual substrate specificity compared with other GH3 enzymes. Xylopyranosyl branching at the non-reducing ends of xyloglucan oligosaccharides is vital for IpeA activity. We solved the crystal structure of IpeA with isoprimeverose at 2.4?Å resolution, showing that the structure of IpeA formed a dimer and was composed of three domains: an N-terminal (β/α)8 TIM-barrel domain, α/β/α sandwich fold domain, and a C-terminal fibronectin-like domain. The catalytic TIM-barrel domain possessed a catalytic nucleophile (Asp300) and acid/base (Glu524) residues. Interestingly, we found that the cavity of the active site of IpeA was larger than that of other GH3 enzymes, and subsite ?1′ played an important role in its activity. The glucopyranosyl and xylopyranosyl residues of isoprimeverose were located at subsites ?1 and ?1′, respectively. Gln58 and Tyr89 contributed to the interaction with the xylopyranosyl residue of isoprimeverose through hydrogen bonding and stacking effects, respectively. Our findings provide new insights into the substrate recognition of GH3 enzymes.  相似文献   

19.
Dibenzothiophene (DBT) is a typical sulfur‐containing compound found in fossil fuels. This compound and its derivatives are resistant to the hydrodesulfurization method often used in industry, but they are susceptible to enzymatic desulfurization via the 4S pathway, which is a well‐studied biochemical pathway consisting of four enzymes. DBT monooxygenase (DszC) from Rhodococcus erythropolis is involved in the first step of the 4S pathway. We determined the crystal structure of DszC, which reveals that, in contrast to several homologous proteins, the C‐terminus (410–417) of DszC participates in the stabilization of the substrate‐binding pocket. Analytical ultracentrifugation analysis and enzymatic assays confirmed that the C‐terminus is important for the stabilization of the active conformation of the substrate‐binding pocket and the tetrameric state. Therefore, the C‐terminus of DszC plays a significant role in the catalytic activity of this enzyme. Proteins 2014; 82:2733–2743. © 2014 Wiley Periodicals, Inc.  相似文献   

20.
Reassessment of the catalytic mechanism of glycogen debranching enzyme   总被引:1,自引:0,他引:1  
W Liu  N B Madsen  C Braun  S G Withers 《Biochemistry》1991,30(5):1419-1424
The amylo-1,6-glucosidase catalytic activity of glycogen debranching enzyme allows it to hydrolyze alpha-D-glucosyl fluoride, in the absence or presence of glycogen or oligosaccharides, releasing equal amounts of fluoride and glucose at rates comparable to those seen with the natural substrates. 2-Deoxy-2-fluoro-alpha-D-glucosyl fluoride is found to be a poor substrate, rather than the covalent inhibitor that would be expected for a glucosidase which catalyzes hydrolysis of the glycosidic linkage with retention of anomeric configuration. In fact, analysis of the glucosidase reaction by NMR reveals that the debranching enzyme hydrolyzes the glycosidic linkage with inversion of configuration, releasing beta-D-glucose from both alpha-glucosyl fluoride and its natural substrate, the phosphorylase limit dextrin. In contrast, its transferase activity necessarily proceeds with retention of configuration. As has been seen with other "inverting" glycosidases, the debranching enzyme releases beta-D-glucose from beta-D-glucosyl fluoride in the presence of oligosaccharides such as maltohexaose and cyclomaltoheptaose but, unlike the others, not in their absence. An intermediate glucosyl-alpha-(1,6)-cyclomaltoheptaose has been detected by NMR analysis. In the presence of a water-soluble carbodiimide, a single mole of glycine ethyl ester is incorporated into each mole of the debranching enzyme, resulting in its inactivation when measured by the combined assay for both transferase and glucosidase activities. Measurement of the latter two activities independently indicates that it is the transferase activity which is inactivated, while the glucosidase activity, measured with alpha-D-glucosyl fluoride as substrate, is unaffected.(ABSTRACT TRUNCATED AT 250 WORDS)  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