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 15 毫秒
1.
BackgroundTransglycosylation (TG) activity is a property of glycosyl hydrolases (GHs) with which new glycosidic bonds are introduced between donor and acceptor sugar molecules. This special property of the GHs has potential to generate longer chain chitooligosaccharides (CHOS) that show elicitor activity in plants. We hypothesize that TG activity could be improved by retaining the substrate for a longer duration in the catalytic site.MethodsFour variants of chitinase D from Serratia proteamaculans (SpChiD) i.e. G119S, G119W, W120A and G201W were analyzed in detail for improved TG activity using high performance liquid chromatography (HPLC) and high resolution mass spectrometry (HRMS). The results were strongly supported by 50 ns molecular dynamics (MD) simulations and estimated solvated interaction energies (SIE).ResultsThe mutant G119W lost much of both hydrolytic and TG activities, while the mutant G201W displayed increased TG. The trajectory of MD simulations of the mutant G119W showed that the indole rings of two adjacent Trp residues create a major hindrance for the DP4 movement towards the catalytic center. Increased van der Waals (vdW) and coulombic interactions between DP4 substrate and the Trp-201 resulted in enhanced TG activity with the mutant G201W. The average number of hydrogen bonds observed for the DP4 substrate was increased for the mutants G119W and G201W compared to SpChiD.ConclusionThe increase in TG activity could be due to partial blocking of product exit of SpChiD.General significanceThis new approach can be used for generating mutants of GHs with improved TG activity to produce longer chain oligosaccharides.  相似文献   

2.
Chitin is an abundant renewable polysaccharide, next only to cellulose. Chitinases are important for effective utilization of this biopolymer. Chitinase D from Serratia proteamaculans (SpChiD) is a single domain chitinase with both hydrolytic and transglycosylation (TG) activities. SpChiD had less of hydrolytic activity on insoluble polymeric chitin substrates due to the absence of auxiliary binding domains. We improved catalytic efficiency of SpChiD in degradation of insoluble chitin substrates by fusing with auxiliary domains like polycystic kidney disease (PKD) domain and chitin binding protein 21 (CBP21). Of the six different SpChiD fusion chimeras, two C-terminal fusions viz. ChiD+PKD and ChiD+CBP resulted in improved hydrolytic activity on α- and β-chitin, respectively. Time-course degradation of colloidal chitin also confirmed that these two C-terminal SpChiD fusion chimeras were more active than other chimeras. More TG products were produced for a longer duration by the fusion chimeras ChiD+PKD and PKD+ChiD+CBP.  相似文献   

3.
Chitinases are known to hydrolyze chitin polymers into smaller chitooligosaccharides. Chitinase from bacterium Serratia proteamaculans (SpChiD) is found to exhibit both hydrolysis and transglycosylation activities. SpChiD belongs to family 18 of glycosyl hydrolases (GH-18). The recombinant SpChiD was crystallized and its three-dimensional structure was determined at 1.49 Å resolution. The structure was refined to an R-factor of 16.2%. SpChiD consists of 406 amino acid residues. The polypeptide chain of SpChiD adopts a (β/α)8 triosephosphate isomerase (TIM) barrel structure. SpChiD contains three acidic residues, Asp149, Asp151 and Glu153 as part of its catalytic scheme. While both Asp149 and Glu153 adopt single conformations, Asp151 is observed in two conformations. The substrate binding cleft is partially obstructed by a protruding loop, Asn30 - Asp42 causing a considerable reduction in the number of available subsites in the substrate binding site. The positioning of loop, Asn30 - Asp42 appears to be responsible for the transglycosylation activity. The structure determination indicated the presence of sulfone Met89 (SMet89). The sulfone methionine residue is located on the surface of the protein at a site where extra domain is attached in other chitinases. This is the first structure of a single domain chitinase with hydrolytic and transglycosylation activities.  相似文献   

4.
Amylomaltase from Thermus aquaticus catalyzes intramolecular transglycosylation of α-1,4 glucans to produce cyclic α-1,4 glucans (cycloamyloses) with degrees of polymerization of 22 and higher. Although the amylomaltase mainly catalyzes the transglycosylation reaction, it also has weak hydrolytic activity, which results in a reduction in the yield of the cycloamyloses. In order to obtain amylomaltase with less hydrolytic activity, random mutagenesis was perfromed for the enzyme gene. Tyr54 (Y54) was identified as the amino acid involved in the hydrolytic activity of the enzyme. When Y54 was replaced with all other amino acids by site-directed mutagenesis, the hydrolytic activities of the mutated enzymes were drastically altered. The hydrolytic activities of the Y54G, Y54P, Y54T, and Y54W mutated enzymes were remarkably reduced compared with that of the wild-type enzyme, while those of the Y54F and Y54K mutated enzymes were similar to that of the wild-type enzyme. Introducing an amino acid replacement at Y54 also significantly affected the cyclization activity of the amylomaltase. The Y54A, Y54L, Y54R, and Y54S mutated enzymes exhibited cyclization activity that was approximately twofold higher than that of the wild-type enzyme. When the Y54G mutated enzyme was employed for cycloamylose production, the yield of cycloamyloses was more than 90%, and there was no decrease until the end of the reaction.  相似文献   

5.
Transglycosylation (TG) by family 18 chitinases is of special interest due to the many biological applications of long-chain chitooligosaccharides (CHOS). In the current study, the TG activity of chitinase A from Stenotrophomonas maltophilia (StmChiA) was improved through structure-guided mutations within and around the active site. Three independent mutants were created, targeting Trp residues from the − 3 and − 1 subsites and the central catalytic Asp from the DxDxE motif of StmChiA. The former was replaced with Ala and the latter with Asn. Changes in the hydrolytic and TG activities of the enzymes were assessed by monitoring the product profile of each mutant by high-performance liquid chromatography. All three mutants showed increased TG activity. Increased in the higher TG activity of mutant W306A was accompanied by increased hydrolysis. However, this mutant also accumulated substantial amounts of TG products during the first 15–30 min of the reaction. In contrast, mutants D464N and W679A showed reduced hydrolysis, which was accompanied by the gradual accumulation of TG products up to 12 h. Molecular docking studies with chitohexaose showed that the side chains of Trp residues mediate stacking interactions with sugar residues from the − 3 and − 1 subsites, indicating the importance of these residues in the enzymatic activity of StmChiA. Overall, mutants of the glycon-binding site (W306A and W679A) appear to produce long-chain CHOS more efficiently than the catalytic mutant D464N.  相似文献   

6.
The mutations (G228S, A391T and F439W) and duplication of the acetylcholinesterase (AChE) gene (Tuace) are involved in monocrotophos resistance in the two-spotted spider mites, Tetranychus urticae (Kwon et al., 2010a, Kwon et al., 2010b). The overexpression of T. urticae AChE (TuAChE) as a result of Tuace duplication was confirmed in several field-collected populations by Western blotting using an AChE-specific antibody. To investigate the effects of each mutation on the insensitivity and fitness cost of AChE, eight variants of TuAChE were expressed in vitro using the baculovirus expression system. Kinetic analysis revealed that the G228S and F439W mutations confer approximately 26-fold and 99-fold increases in the insensitivity to monocrotophos, respectively, whereas the insensitivity increased over 1165-fold in the AChE with double mutations. Nevertheless, the presence of these mutations reduced the catalytic efficiency of AChE significantly. In particular, the TuAChE having both mutations together exhibited a 17.8~27.1-fold reduced catalytic efficiency, suggesting an apparent fitness cost in the monocrotophos-resistant mites. The A391T mutation did not change the kinetic properties of either the substrate or inhibitor when present alone but mitigated the negative impacts of the F439 mutation. To simulate the catalytic activity of the overexpressed TuAChE in two T. urticae strains (approximately 6 copies for AD strain vs. 2 copies for PyriF strain), appropriate TuAChE variants were combined to make up the desired AChE copies and mutation frequencies, and their enzyme kinetics were determined. The reconstituted 6-copy and 2-copy TuAChEs exhibited catalytic efficiency levels comparable to those of a single-copy wildtype TuAChE, suggesting that, if mutations are present, multiple copies of AChE are required to restore a normal level of catalytic activity in the monocrotophos-resistant mites. In summary, the present study provides clear evidence that Tuace duplication resulted in the proportional overexpression of AChE, which was necessary to compensate for the reduced catalytic activity of AChE caused by mutations.  相似文献   

7.
Point mutations of the active-site residues Trp168, Tyr171, Trp275, Trp397, Trp570 and Asp392 were introduced to Vibrio carchariae chitinase A. The modeled 3D structure of the enzyme illustrated that these residues fully occupied the substrate binding cleft and it was found that their mutation greatly reduced the hydrolyzing activity against pNP-[GlcNAc]2 and colloidal chitin. Mutant W397F was the only exception, as it instead enhanced the hydrolysis of the pNP substrate to 142% and gave no activity loss towards colloidal chitin. The kinetic study with the pNP substrate demonstrated that the mutations caused impaired Km and kcat values of the enzyme. A chitin binding assay showed that mutations of the aromatic residues did not change the binding equilibrium. Product analysis by thin layer chromatography showed higher efficiency of W275G and W397F in G4–G6 hydrolysis over the wild type enzyme. Though the time course of colloidal chitin hydrolysis displayed no difference in the cleavage behavior of the chitinase variants, the time course of G6 hydrolysis exhibited distinct hydrolytic patterns between wild-type and mutants W275G and W397F. Wild type initially hydrolyzed G6 to G4 and G2, and finally G2 was formed as the major end product. W275G primarily created G2–G5 intermediates, and later G2 and G3 were formed as stable products. In contrast, W397F initially produced G1–G5, and then the high-Mr intermediates (G3–G5) were broken down to G1 and G2 end products. This modification of the cleavage patterns of chitooligomers suggested that residues Trp275 and Trp397 are involved in defining the binding selectivity of the enzyme to soluble substrates.  相似文献   

8.
Xylanases produce xylooligosaccharides from xylan and have thus attracted increasing attention for their usefulness in industrial applications. Previously, we demonstrated that the GH11 xylanase XynLC9 from Bacillus subtilis formed xylobiose and xylotriose as the major products with negligible production of xylose when digesting corncob-extracted xylan. Here, we aimed to improve the catalytic performance of XynLC9 via protein engineering. Based on the sequence and structural comparisons of XynLC9 with the xylanases Xyn2 from Trichoderma reesei and Xyn11A from Thermobifida fusca, we identified the N-terminal residues 5-YWQN-8 in XynLC9 as engineering hotspots and subjected this sequence to site saturation and iterative mutagenesis. The mutants W6F/Q7H and N8Y possessed a 2.6- and 1.8-fold higher catalytic activity than XynLC9, respectively, and both mutants were also more thermostable. Kinetic measurements suggested that W6F/Q7H and N8Y had lower substrate affinity, but a higher turnover rate (kcat), which resulted in increased catalytic efficiency than WT XynLC9. Furthermore, the W6F/Q7H mutant displayed a 160% increase in the yield of xylooligosaccharides from corncob-extracted xylan. Molecular dynamics simulations revealed that the W6F/Q7H and N8Y mutations led to an enlarged volume and surface area of the active site cleft, which provided more space for substrate entry and product release and thus accelerated the catalytic activity of the enzyme. The molecular evolution approach adopted in this study provides the design of a library of sequences that captures functional diversity in a limited number of protein variants.  相似文献   

9.
In order to get insight into the origin of apparent negative cooperativity observed for F1-ATPase, we compared ATPase activity and ATPMg binding of mutant subcomplexes of thermophilic F1-ATPase, α(W463F)3β(Y341W)3γ and α(K175A/T176A/W463F)3β(Y341W)3γ. For α(W463F)3β(Y341W)3γ, apparent Km's of ATPase kinetics (4.0 and 233 μM) did not agree with apparent Km's deduced from fluorescence quenching of the introduced tryptophan residue (on the order of nM, 0.016 and 13 μM). On the other hand, in case of α(K175A/T176A/W463F)3β(Y341W)3γ, which lacks noncatalytic nucleotide binding sites, the apparent Km of ATPase activity (10 μM) roughly agreed with the highest Km of fluorescence measurements (27 μM). The results indicate that in case of α(W463F)3β(Y341W)3γ, the activating effect of ATP binding to noncatalytic sites dominates overall ATPase kinetics and the highest apparent Km of ATPase activity does not represent the ATP binding to a catalytic site. In case of α(K175A/T176A/W463F)3β(Y341W)3γ, the Km of ATPase activity reflects the ATP binding to a catalytic site due to the lack of noncatalytic sites. The Eadie-Hofstee plot of ATPase reaction by α(K175A/T176A/W463F)3β(Y341W)3γ was rather linear compared with that of α(W463F)3β(Y341W)3γ, if not perfectly straight, indicating that the apparent negative cooperativity observed for wild-type F1-ATPase is due to the ATP binding to catalytic sites and noncatalytic sites. Thus, the frequently observed Km's of 100-300 μM and 1-30 μM range for wild-type F1-ATPase correspond to ATP binding to a noncatalytic site and catalytic site, respectively.  相似文献   

10.
To investigate the roles of the active site residues in the catalysis of Bacillus thuringiensis WB7 chitinase, twelve mutants, F201L, F201Y, G203A, G203D, D205E, D205N, D207E, D207N, W208C, W208R, E209D and E209Q were constructed by site-directed mutagenesis. The results showed that the mutants F201L, G203D, D205N, D207E, D207N, W208C and E209D were devoid of activity, and the loss of the enzymatic activities for F201Y, G203A, D205E, W208R and E209Q were 72, 70, 48, 31 and 29%, respectively. The pH-activity profiles indicated that the optimum pH for the mutants as well as for the wildtype enzyme was 8.0. E209Q exhibited a broader active pH range while D205E, G203A and F201Y resulted in a narrower active pH range. The pH range of activity reduced 1 unit for D205E, and 2 units for G203A and F201Y. The temperature-activity profiles showed that the optimum temperature for other mutants as well as wildtype enzyme was 60°C, but 50°C for G203A, which suggested that G203A resulted in a reduction of thermostability. The study indicated that the six active site residues involving in mutagenesis played an important part in WB7 chitinase. In addition, the catalytic mechanisms of the six active site residues in WB7 chitinase were discussed.  相似文献   

11.
Mutations of the tryptophan residues in the tryptophan-track of the N-terminal domain (W33F/Y and W69F/Y) and in the catalytic domain (W245F/Y) of Serratia sp. TU09 Chitinase 60 (CHI60) were constructed, as single and double point substitutions to either phenylalanine or tyrosine. The enzyme-substrate interaction and mode of catalysis, exo/endo-type, of wild type CHI60 and mutant enzymes on soluble (partially N-acetylated chitin), amorphous (colloidal chitin), and crystalline (β-chitin) substrates were studied. All CHI60 mutants exhibited a reduced substrate binding activity on colloidal chitin. CHI60 possesses a dual mode of catalysis with both exo- and endo-type activities allowing the enzyme to work efficiently on various substrate types. CHI60 preferentially uses the endo-type mode on soluble and amorphous substrates and the exo-type mode on crystalline substrate. However, the prevalent mode of hydrolysis mediated by CHI60 is regulated by ionic strength. Slightly elevated ionic strength, 0.1-0.2 M NaCl, which promotes enzyme-substrate interactions, enhances CHI60 hydrolytic activity on amorphous substrate and, interestingly, on partially N-acetylated chitin. High ionic strength, 0.5-2.0 M NaCl, prevents the enzyme from dissociating from amorphous substrate, occupying the enzyme in an enzyme-substrate non-productive complex. However, on crystalline substrates, the activity of CHI60 was only inhibited approximately 50% at high ionic strength, suggesting that the enzyme hydrolyzes crystalline substrates with an exo-type mode processively while remaining tightly bound to the substrate. Moreover, substitution of Trp-33 to either phenylalanine or tyrosine reduced the activity of the enzyme at high ionic strength, suggesting an important role of Trp-33 on enzyme processivity.  相似文献   

12.
Oligopeptidase A (OpdA) belongs to the M3A subfamily of bacterial peptidases with catalytic and structural properties similar to mammalian thimet-oligopeptidase (TOP) and neurolysin (NEL). The three enzymes have four conserved Tyr residues on a flexible loop in close proximity to the catalytic site. In OpdA, the flexible loop is formed by residues 600-614 (600SHIFAGGYAAGYYSY614). Modeling studies indicated that in OpdA the Tyr607 residue might be involved in the recognition of the substrate with a key role in catalysis. Two mutants were constructed replacing Tyr607 by Phe (Y607F) or Ala (Y607A) and the influence of the site-directed mutagenesis in the catalytic process was examined. The hydrolysis of Abz-GXSPFRQ-EDDnp derivatives (Abz = ortho-aminobenzoic acid; EDDnp N-[2,4-dinitrophenyl]-ethylenediamine; X = different amino acids) was studied to compare the activities of wild-type OpdA (OpdA WT) and those of Y607F and Y607A mutants The results indicated that OpdA WT cleaved all the peptides only on the X-S bond whereas the Y607F and Y607A mutants were able to hydrolyze both the X-S and the P-F bonds. The kinetic parameters showed the importance of Tyr607 in OpdA catalytic activity as its substitution promoted a decrease in the kcat/Km value of about 100-fold with Y607F mutant and 1000-fold with Y607A. Both mutations, however, did not affect protein folding as indicated by CD and intrinsic fluorescence analysis. Our results indicate that the OpdA Tyr607 residue plays an important role in the enzyme-substrate interaction and in the hydrolytic activity.  相似文献   

13.
Amylomaltase from Thermus aquaticus catalyzes intramolecular transglycosylation of alpha-1,4 glucans to produce cyclic alpha-1,4 glucans (cycloamyloses) with degrees of polymerization of 22 and higher. Although the amylomaltase mainly catalyzes the transglycosylation reaction, it also has weak hydrolytic activity, which results in a reduction in the yield of the cycloamyloses. In order to obtain amylomaltase with less hydrolytic activity, random mutagenesis was perfromed for the enzyme gene. Tyr54 (Y54) was identified as the amino acid involved in the hydrolytic activity of the enzyme. When Y54 was replaced with all other amino acids by site-directed mutagenesis, the hydrolytic activities of the mutated enzymes were drastically altered. The hydrolytic activities of the Y54G, Y54P, Y54T, and Y54W mutated enzymes were remarkably reduced compared with that of the wild-type enzyme, while those of the Y54F and Y54K mutated enzymes were similar to that of the wild-type enzyme. Introducing an amino acid replacement at Y54 also significantly affected the cyclization activity of the amylomaltase. The Y54A, Y54L, Y54R, and Y54S mutated enzymes exhibited cyclization activity that was approximately twofold higher than that of the wild-type enzyme. When the Y54G mutated enzyme was employed for cycloamylose production, the yield of cycloamyloses was more than 90%, and there was no decrease until the end of the reaction.  相似文献   

14.
15.
Thermobifida fusca cutinase (TfCut2) is a carboxylesterase (CE) which degrades polyethylene terephthalate (PET) as well as its degradation intermediates [such as oligoethylene terephthalate (OET), or bis-/mono-hydroxyethyl terephthalate (BHET/MHET)] into terephthalic acid (TPA). Comparisons of the surfaces of certain CEs (including TfCut2) were combined with docking and molecular dynamics simulations involving 2HE-(MHET)3, a three-terephthalate OET, to support the rational design of 22 variants with potential for improved generation of TPA from PET, comprising 15 single mutants (D12L, E47F, G62A, L90A, L90F, H129W, W155F, ΔV164, A173C, H184A, H184S, F209S, F209I, F249A, and F249R), 6 double mutants [H129W/T136S, A173C/A206C, A173C/A210C, G62A/L90F, G62A/F209I, and G62A/F249R], and 1 triple mutant [G62A/F209I/F249R]. Of these, nine displayed no activity, three displayed decreased activity, three displayed comparable activity, and seven displayed increased (~1.3- to ~7.2-fold) activity against solid PET, while all variants displayed activity against BHET. Of the variants that displayed increased activity against PET, four displayed more activity than G62A, the most-active mutant of TfCut2 known till date. Of these four, three displayed even more activity than LCC (G62A/F209I, G62A/F249R, and G62A/F209I/F249R), a CE known to be ~5-fold more active than wild-type TfCut2. These improvements derived from changes in PET binding and not changes in catalytic efficiency.  相似文献   

16.
Proteins of the Ard family are specific inhibitors of type I restriction-modification enzymes. The ArdA of R64 is highly homologous to ColIb-P9 ArdA, differing only by four amino acid residues of the overall 166. However, unlike ColIb-P9 ArdA, which inhibits both the endonuclease and the methylase activities of EcoKI, the R64 ArdA protein inhibits only the endonuclease activity of this enzyme. The mutant forms of R64 ArdA—A29T, S43A, and Y75W, capable of partially reversing the protein to ColIb-P9 ArdA form—were produced by directed mutagenesis. It was demonstrated that only Y75W mutation of these three variants essentially influenced the functional activity of ArdA: the antimodification activity was restored to approximately 90%. It is assumed that R64 ArdA inhibits formation of the complex between unmodified DNA and the R subunit of the type I restriction-modification enzyme EcoKI (R2M2S), which translocates and cleaves DNA. ColIb-P9 ArdA protein is capable of forming the DNA complex not only with the R subunit, but also with the S subunit, which contacts sK site (containing modified adenine residues) in DNA. ArdA bound to the specific sK site inhibits concurrently the endonuclease and methylase activities of EcoKI (R2M2S), while ArdA bound to the nonspecific site in the R subunit blocks only its endonuclease activity.  相似文献   

17.
Butyrylcholinesterase is a serine esterase, closely related to acetylcholinesterase. Both enzymes employ a catalytic triad mechanism for catalysis, similar to that used by serine proteases such as alpha-chymotrypsin. Enzymes of this type are generally considered to be inactive at pH values below 5, because the histidine member of the catalytic triad becomes protonated. We have found that butyrylcholinesterase retains activity at pH 相似文献   

18.
19.
20.
The conserved sequence motif "RxY(T)(S)xx(S)(N)" coordinates flavin binding in NADH:cytochrome b(5) reductase (cb(5)r) and other members of the flavin transhydrogenase superfamily of oxidoreductases. To investigate the roles of Y93, the third and only aromatic residue of the "RxY(T)(S)xx(S)(N)" motif, that stacks against the si-face of the flavin isoalloxazine ring, and P92, the second residue in the motif that is also in close proximity to the FAD moiety, a series of rat cb(5)r variants were produced with substitutions at either P92 or Y93, respectively. The proline mutants P92A, G, and S together with the tyrosine mutants Y93A, D, F, H, S, and W were recombinantly expressed in E. coli and purified to homogeneity. Each mutant protein was found to bind FAD in a 1:1 cofactor:protein stoichiometry while UV CD spectra suggested similar secondary structure organization among all nine variants. The tyrosine variants Y93A, D, F, H, and S exhibited varying degrees of blue-shift in the flavin visible absorption maxima while visible CD spectra of the Y93A, D, H, S, and W mutants exhibited similar blue-shifted maxima together with changes in absorption intensity. Intrinsic flavin fluorescence was quenched in the wild type, P92S and A, and Y93H and W mutants while Y93A, D, F, and S mutants exhibited increased fluorescence when compared to free FAD. The tyrosine variants Y93A, D, F, and S also exhibited greater thermolability of FAD binding. The specificity constant (k(cat)/K(m)(NADH)) for NADH:FR activity decreased in the order wild type > P92S > P92A > P92G > Y93F > Y93S > Y93A > Y93D > Y93H > Y93W with the Y93W variant retaining only 0.5% of wild-type efficiency. Both K(s)(H4NAD) and K(s)(NAD+) values suggested that Y93A, F, and W mutants had compromised NADH and NAD(+) binding. Thermodynamic measurements of the midpoint potential (E degrees ', n = 2) of the FAD/FADH(2) redox couple revealed that the potentials of the Y93A and S variants were approximately 30 mV more positive than that of wild-type cb(5)r (E degrees ' = -268 mV) while that of Y93H was approximately 30 mV more negative. These results indicate that neither P92 nor Y93 are critical for flavin incorporation in cb(5)r and that an aromatic side chain is not essential at position 93, but they demonstrate that Y93 forms contacts with the FAD that effectively modulate the spectroscopic, catalytic, and thermodynamic properties of the bound cofactor.  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