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.
The lysozyme-catalyzed reaction of chitooligosaccharide was carried out in a continuous flow system in which the solution of substrate, chitooligosaccharide [(GlcNAc)n], flowed into the lysozyme solution in an ultrafiltration apparatus and the products were filtered off. The filtrate was continuously collected in test tubes with the aid of a fraction collector. The product distribution in each fraction was analyzed by high performance gel filtration. Using (GlcNAc)5 as the substrate, the concentrations of products, (GlcNAc)1----4, increased gradually and came to the steady state when the volume of the outflow amounted to sixfold of the inside volume. Before reaching the steady state, the product distribution was quite different from that observed in the closed reaction system, in which the reaction species are not exchangeable through the boundary of the system. The outflows of (GlcNAc)3-5 were delayed in comparison with those of GlcNAc and (GlcNAc)2. The delay period increased with the decrease in substrate concentration, and was shortened by using the [Asp 101 or Trp 62]-modified lysozyme instead of the native lysozyme. These results suggest that the delay in the (GlcNAc)3-5 outflows is caused by the nonproductive binding of the oligosaccharide to the lysozyme molecule. The profile of the flow reaction yields information not only on the catalytic efficiency but also on the substrate binding efficiency of the lysozyme.  相似文献   

2.
In the three-dimensional structure of a rice class I chitinase (OsChia1b) determined recently, a loop structure (loop II) is located at the end of the substrate-binding cleft, and is thus suggested to be involved in substrate binding. In order to test this assumption, deletion of the loop II region from the catalytic domain of OsChia1b and replacement of Trp159 in loop II with Ala were carried out. The loop II deletion and the W159A mutation increased hydrolytic activity not only towards (GlcNAc)6 but also towards polysaccharide substrates. Similar results were obtained for kcat/Km values determined for substrate reduced-(GlcNAc)5. The two mutations shifted the splitting positions in (GlcNAc)6 to the reducing end side, but the shift was less intensive in the Trp mutant. Theoretical analysis of the reaction time course indicated that sugar residue affinity at the +3 subsite was reduced from -2 kcal/mol to +0.5 kcal/mol by loop II deletion. Reduced affinity at the +3 subsite might enhance the release of product fragments, resulting in higher turnover and higher enzymatic activities. Thus, we concluded that loop II is involved in sugar residue binding at the +3 subsite, but that Trp159 itself appears to contribute only partly to sugar residue interaction at the subsite.  相似文献   

3.
The stacking interaction between a tyrosine residue and the sugar ring at the catalytic subsite -1 is strictly conserved in the glycoside hydrolase family 13 enzymes. Replacing Tyr100 with leucine in cyclodextrin glycosyltransferase (CGTase) from Bacillus sp. 1011 to prevent stacking significantly decreased all CGTase activities. The adjacent stacking interaction with both Phe183 and Phe259 onto the sugar ring at subsite +2 is essentially conserved among CGTases. F183L/F259L mutant CGTase affects donor substrate binding and/or acceptor binding during transglycosylation [Nakamura et al. (1994) Biochemistry 33, 9929-9936]. To elucidate the precise role of carbohydrate/aromatic stacking interaction at subsites -1 and +2 on the substrate binding of CGTases, we analyzed the X-ray structures of wild-type (2.0 A resolution), and Y100L (2.2 A resolution) and F183L/F259L mutant (1.9 A resolution) CGTases complexed with the inhibitor, acarbose. The refined structures revealed that acarbose molecules bound to the Y100L mutant moved from the active center toward the side chain of Tyr195, and the hydrogen bonding and hydrophobic interaction between acarbose and subsites significantly diminished. The position of pseudo-tetrasaccharide binding in the F183L/F259L mutant was closer to the non-reducing end, and the torsion angles of glycosidic linkages at subsites -1 to +1 on molecule 1 and subsites -2 to -1 on molecule 2 significantly changed compared with that of each molecule of wild-type-acarbose complex to adopt the structural change of subsite +2. These structural and biochemical data suggest that substrate binding in the active site of CGTase is critically affected by the carbohydrate/aromatic stacking interaction with Tyr100 at the catalytic subsite -1 and that this effect is likely a result of cooperation between Tyr100 and Phe259 through stacking interaction with substrate at subsite +2.  相似文献   

4.
The interactions of the substrate analogues, GlcNAc, beta-methyl GlcNAc, (GlcNAc)2, and (GlcNAc)3, with turkey egg-white lysozyme [ED 3.2.1.17], in which the Asp 101 of hen lysozyme is replaced by Gly, were studied at various pH values by measuring changes in the circular dichroic (CD) band at 295 nm. Results were compared with those for hen egg-white lysozyme. The modes of binding of these substrate analogues to turkey lysozyme were very similar to those hen lysozyme except for the participation of Asp 101 in hen lysozyme. The ionization constants of the catalytic carboxyls, Glu 35 and Asp 52, in the turkey lysozyme-(GlcNAc)3 complex were determined by measuring the pH dependence of the CD band at 304 nm, which originates from Trp 108 near the catalytic carboxyls. The ionization behavior of the catalytic carboxyls of turkey lysozyme in the presence and absence of (GlcNAc)3 was essentially the same as that for hen lysozyme. The pH dependence of the binding constant of (GlcNAc)3 to hen lysozyme was compared with that to turkey lysozyme between pH 2 and 8. The pH dependence of the binding constant for (GlcNAc)3 to turkey lysozyme could be interpreted entirely in terms of perturbation of catalytic carboxyls. In the case of hen lysozyme, it was interpreted in terms of perturbation of the catalytic carboxyls and Asp 101 in the substrate-binding site. The pK values of Asp 101 in hen lysozyme and the hen lysozyme-(GLcNAc)3 complex were 4.5 and 3.4, respectively. The binding constants of (GlcNAc)3 to lysozyme molecules with different microscopic protonation forms, with respect to the catalytic carboxyls, were estimated. The binding constant of lysozyme, in which Asp 52 and Glu 35 are deprotonated, to (GlcNAc)3 was the smallest. The other three species had similar binding constant to (GlcNAc)3.  相似文献   

5.
The structure of the complex between a catalytically compromised family 10 xylanase and a xylopentaose substrate has been determined by X-ray crystallography and refined to 3.2 A resolution. The substrate binds at the C-terminal end of the eightfold betaalpha-barrel of Pseudomonas fluorescens subsp. cellulosa xylanase A and occupies substrate binding subsites -1 to +4. Crystal contacts are shown to prevent the expected mode of binding from subsite -2 to +3, because of steric hindrance to subsite -2. The loss of accessible surface at individual subsites on binding of xylopentaose parallels well previously reported experimental measurements of individual subsites binding energies, decreasing going from subsite +2 to +4. Nine conserved residues contribute to subsite -1, including three tryptophan residues forming an aromatic cage around the xylosyl residue at this subsite. One of these, Trp 313, is the single residue contributing most lost accessible surface to subsite -1, and goes from a highly mobile to a well-defined conformation on binding of the substrate. A comparison of xylanase A with C. fimi CEX around the +1 subsite suggests that a flatter and less polar surface is responsible for the better catalytic properties of CEX on aryl substrates. The view of catalysis that emerges from combining this with previously published work is the following: (1) xylan is recognized and bound by the xylanase as a left-handed threefold helix; (2) the xylosyl residue at subsite -1 is distorted and pulled down toward the catalytic residues, and the glycosidic bond is strained and broken to form the enzyme-substrate covalent intermediate; (3) the intermediate is attacked by an activated water molecule, following the classic retaining glycosyl hydrolase mechanism.  相似文献   

6.
The reaction pattern of an extracellular chitin deacetylase from a Deuteromycete, Colletotrichum lindemuthianum ATCC 56676, was investigated by use of chitooligosaccharides [(GlcNAc)(n)(), n = 3-6] and partially N-deacetylated chitooligosaccharides as substrates. When 0.5% of (GlcNAc)(n)() was deacetylated, the corresponding monodeacetylated products were initially detected without any processivity, suggesting the involvement of a multiple-chain mechanism for the deacetylation reaction. The structural analysis of these first-step products indicated that the chitin deacetylase strongly recognizes a sequence of four N-acetyl-D-glucosamine (GlcNAc) residues of the substrate (the subsites for the four GlcNAc residues are defined as -2, -1, 0, and +1, respectively, from the nonreducing end to the reducing end), and the N-acetyl group in the GlcNAc residue positioned at subsite 0 is exclusively deacetylated. When substrates of a low concentration (100 microM) were deacetylated, the initial deacetylation rate for (GlcNAc)(4) was comparable to that of (GlcNAc)(5), while deacetylation of (GlcNAc)(3) could not be detected. Reaction rate analyses of partially N-deacetylated chitooligosaccharides suggested that subsite -2 strongly recognizes the N-acetyl group of the GlcNAc residue of the substrate, while the deacetylation rate was not affected when either subsite -1 or +1 was occupied with a D-glucosamine residue instead of GlcNAc residue. Thus, the reaction pattern of the chitin deacetylase is completely distinct from that of a Zygomycete, Mucor rouxii, which produces a chitin deacetylase for accumulation of chitosan in its cell wall.  相似文献   

7.
Rhea lysozyme was analyzed for its enzymatic properties both lytic and oligomer activities to reveal the structural and functional relationships of goose type lysozyme. Rhea lysozyme had the highest lytic activity at pH 6, followed by ostrich and goose at pH 5.5-6, whereas the optimum of cassowary was at pH 5. pH profile was correlated to the net charge of each molecule surface. On the other hand, the pH optimum for oligomer substrate was found to be pH 4, indicating the mechanism of rhea catalysis as a general acid. The time-course of the reaction was studied using beta-1,4-linked oligosaccharide of N-acetylglucosamine (GlcNAc) with a polymerization degree of n ((GlcNAc)n) (n=4, 5, and 6) as the substrate. This enzyme hydrolyzed (GlcNAc)6 in an endo-splitting manner, which produced (GlcNAc)3+(GlcNAc)3 predominating over that to (GlcNAc)2+ (GlcNAc)4. This indicates that the lysozyme hydrolyzed preferentially the third glycosidic linkage from the nonreducing end. Theoretical analysis has shown the highest rate constant value at 1.5 s-1 with (GlcNAc)6. This confirmed six substrate binding subsites as goose lysozyme (Honda, Y., and Fukamizo, T., Biochim. Biophys. Acta, 1388, 53-65 (1998)). The different binding free energy values for subsites B, C, F, and G from goose lysozyme might responsible for the amino acid substitutions, Asn122Ser and Phe123Met, located at the subsite B.  相似文献   

8.
Hydrolysis of small substrates (maltose, maltotriose and o-nitrophenylmaltoside) catalysed by porcine pancreatic alpha-amylase was studied from a kinetic viewpoint over a wide range of substrate concentrations. Non-linear double-reciprocal plots are obtained at high maltose, maltotriose and o-nitrophenylmaltoside concentrations indicating typical substrate inhibition. These results are consistent with the successive binding of two molecules of substrate per enzyme molecule with dissociation constants Ks1 and Ks2. The Hill plot, log [v/(V-v)] versus log [S], is clearly biphasic and allows the dissociation constants of the ES1 and ES2 complexes to be calculated. Maltose and maltotriose are inhibitors of the amylase-catalysed amylose and o-nitrophenylmaltoside hydrolysis. The inhibition is of the competitive type. The (apparent) inhibition constant Kiapp varies with the inhibitor concentration. These results are also consistent with the successive binding of at least two molecules of maltose or maltotriose per amylase molecule with the dissociation constants Ki1 and Ki2. These inhibition studies show that small substrates and large polymeric ones are hydrolysed at the same catalytic site(s). The values of the dissociation constants Ks1 and Ki1 of the maltose-amylase complexes are identical. According to the five-subsite energy profile previously determined, at low concentration, maltose (as substrate and as inhibitor) binds to the same two sites (4,5) or (3,4), maltotriose (as substrate and as inhibitor) and o-nitrophenyl-maltoside (as substrate) bind to the same three subsites (3,4,5). The dissociation constants Ks2 and Ki2 determined at high substrate and inhibitor concentration are consistent with the binding of the second ligand molecule at a single subsite. The binding mode of the second molecule of maltose (substrate) and o-nitrophenylmaltoside remains uncertain, very likely because of the inaccuracy due to simplifications in the calculations of the subsite binding energies. No binding site(s) outside the catalytic one has been taken into account in this model.  相似文献   

9.
M C Shao  G Krudy  P R Rosevear  F Wold 《Biochemistry》1989,28(9):4077-4083
The effect of the protein environment on the reaction sequence and the relative rates of two two-step reactions involved in the biosynthesis of complex glycans in glycoproteins has been explored by comparing the processing of biotinylated substrates either free or bound to avidin. By use of biotinyl and biotinamidohexanoyl derivatives, the display of the glycan in a proximal and distal association with the avidin surface could also be assessed. Mannosidase II removes two Man residues from the substrate GlcNAcMan5GlcNAc2-R to yield GlcNAcMAn3GlcNAc2-R. The NMR spectra of the substrate, intermediate, and product showed that the first Man is removed from the 6-arm of the substrate. The rate constants for the first and second step (estimated by direct analysis of the reactants by anion-exchange chromatography with a pulsed amperometric detector) were determined to be about 0.05 and 0.08 min-1, respectively, for the free substrates. In the proximal complex k1 was reduced 80-fold, and the k2 step could not be observed under the same conditions. In the distal complex both k1 and k2 were reduced about 8-fold. Sialyl transferases transfer Sia from CMP-Sia to the biantennary substrate Gal2GlcNAc2-Man3GlcNA2-R to yield the product Sia2Gal2-GlcNAc2Man3GlcNAc2-R with the Sia linked either 2-3 or 2-6 to the Gal residues. The NMR spectra showed that the first step involved the Gal on the 3-arm of the substrate and that both Sia residues were added 2-6.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

10.
The crystal structure of Pterocarpus angolensis lectin is determined in its ligand-free state, in complex with the fucosylated biantennary complex type decasaccharide NA2F, and in complex with a series of smaller oligosaccharide constituents of NA2F. These results together with thermodynamic binding data indicate that the complete oligosaccharide binding site of the lectin consists of five subsites allowing the specific recognition of the pentasaccharide GlcNAc beta(1-2)Man alpha(1-3)[GlcNAc beta(1-2)Man alpha(1-6)]Man. The mannose on the 1-6 arm occupies the monosaccharide binding site while the GlcNAc residue on this arm occupies a subsite that is almost identical to that of concanavalin A (con A). The core mannose and the GlcNAc beta(1-2)Man moiety on the 1-3 arm on the other hand occupy a series of subsites distinct from those of con A.  相似文献   

11.
Hrmova M  Fincher GB 《Carbohydrate research》2007,342(12-13):1613-1623
Higher plant, family GH3 beta-D-glucan glucohydrolases exhibit exo-hydrolytic and retaining (e-->e) mechanisms of action and catalyze the removal of single glucosyl residues from the non-reducing termini of beta-D-linked glucosidic substrates, with retention of anomeric configuration. The broad specificity beta-D-glucan glucohydrolases are likely to play roles in cell wall re-modelling, turn-over of cell wall components and possibly in plant defence reactions against pathogens. Crystal structures of the barley beta-D-glucan glucohydrolase, obtained from both native enzyme and from the enzyme in complex with a substrate analogues and mechanism-based inhibitors, have enabled the basis of substrate specificity, the mechanism of catalysis, and the role of domain movements during the catalytic cycle to be defined in precise molecular terms. The active site of the enzyme forms a shallow 'pocket' that is located at the interface of two domains of the enzyme and accommodates two glucosyl residues. The propensity of the enzyme to hydrolyze a broad range of substrates with (1-->2)-, (1-->3)-, (1-->4)- and (1-->6)-beta-D-glucosidic linkages is explained from crystal structures of the enzyme in complex with non-hydrolysable S-glycoside substrate analogues, and from molecular modelling. During binding of gluco-oligosaccharides, the glucosyl residue at subsite -1 is locked in a highly constrained position, but the glucosyl residue at the +1 subsite is free to adjust its position between two tryptophan residues positioned at the entry of the active site pocket. The flexibility at subsite +1 and the projection of the remainder of the substrate away from the pocket provide a structural rationale for the capacity of the enzyme to accommodate and hydrolyze glucosides with different linkage positions and hence different overall conformations. While mechanism-based inhibitors with micromolar Ki constants bind in the active site of the enzyme and form esters with the catalytic nucleophile, transition-state mimics bind with their 'glucose' moieties distorted into the 4E conformation, which is critical for the nanomolar binding of these inhibitors to the enzyme. The glucose product of the reaction, which is released from the non-reducing termini of substrates, remains bound to the beta-D-glucan glucohydrolase in the -1 subsite of the active site, until a new substrate molecule approaches the enzyme. If dissociation of the glucose from the enzyme active site could be synchronized throughout the crystal, time-resolved Laue X-ray crystallography could be used to follow the conformational changes that occur as the glucose product diffuses away and the incoming substrate is bound by the enzyme.  相似文献   

12.
The interactions of the substrate analogs beta-methyl-GlcNAc, (GlcNAc)2, and (GlcNAc)3 with hen egg-white lysozyme [EC 3.2.1.17] in which an ester linkage had been formed between Glu 35 and Trp 108 (108 ester lysozyme), were studied by the circular dichroic and fluorescence techniques, and were compared with those for intact lysozyme. The binding constants of beta-methyl-GlcNAc and (GlcNAc)2 to 108 ester lysozyme were essentially the same as those for intact lysozyme in the pH range of 1 to 5. Above pH 5, the binding constants of these saccharides to 108 ester lysozyme did not change with pH, while the binding constants to intact lysozyme decreased. This indicates that Glu 35 (pK 6.0 in intact lysozyme) participates in the binding of these saccharides. The extent and direction of the pK shifts of Asp 52 (pK 3.5), Asp 48 (pK 4.4), and Asp 66 (pK 1.3) observed when beta-methyl-GlcNAc is bound to 108 ester lysozyme were the same as those for intact lysozyme. The participation of Asp 101 and Asp 66 in the binding of (GlcNAc)2 to 108 ester lysozyme was also the same as that for intact lysozyme. These findings indicate that the conformations of subsites B and C are not changed by the formation of the ester linkage. On the other hand, the binding constants of (GlcNAc)3 to 108 ester lysozyme were higher than those for intact lysozyme at all pH values studied. This result is interpreted in terms of an increase in the affinity for a GlcNAc residue of subsite D, which is situated near the esterified Glu 35.  相似文献   

13.
The catalytic mechanism for the mono-ADP-ribosyltransferase activity of Pseudomonas aeruginosa exotoxin A was investigated by steady-state and stopped-flow kinetic analyses. The rate constants for binding of the NAD(+) substrate to the enzyme were found to be 4.7 +/- 0.4 microM(-1) s(-1) and 194 +/- 15 s(-1) for k(on) and k(off), respectively. The k(on) and k(off) rate constants for the eEF-2 substrate binding to the enzyme were 320 +/- 39 microM(-1) s(-1) and 131 +/- 22 s(-1), respectively. A potent, competitive inhibitor against the enzyme, 1,8-naphthalimide, bound the enzyme with k(on) and k(off) rates of 82 +/- 9 microM(-1) s(-1) and 51 +/- 6 s(-1), respectively. Furthermore, the binding on and off rates for the reaction products, ADP-ribose and nicotinamide, were too rapid for detection with the stopped-flow technique. Investigation of the pre-steady-state kinetics for the ADP-ribose transferase activity of the toxin-enzyme showed that there is no pre-steady-state complex formed during the catalytic cycle. Binding of NAD+ and smaller compounds representing the various parts of this substrate were investigated by the fluorescence quenching of the intrinsic toxin fluorescence. The binding data revealed a significant structural change in the enzyme upon NAD+ binding that could not be accounted for on the basis of the sum of the structural changes induced by the various NAD+ constituents. Product inhibition studies were conducted with nicotinamide and eEF-2-ADP-ribose, and the results indicate that the reaction involves a random-order ternary complex mechanism. Detailed kinetic analysis revealed that the eEF-2 substrate shows sigmoidal kinetic behavior with the enzyme, and fluorescence resonance energy transfer measurements indicated that wheat germ eEF-2 is oligomeric in solution.  相似文献   

14.
To understand better the role of subsites E and F in lysozyme-catalyzed reactions, mutant enzymes, in which Arg114, located on the right side of subsites E and F in hen egg-white lysozyme (HEL), was replaced with Lys, His, or Ala, were prepared. Replacement of Arg114 with His or Ala decreased hydrolytic activity toward an artificial substrate, glycol chitin, while replacement with Lys had little effect. Kinetic analysis with the substrate N-acetylglucosamine pentamer, (GlcNAc)(5), revealed that the replacement for the Arg residue reduced the binding free energies of E-F sites and the rate constant of transglycosylation. The rate constant of transglycosylation for R114A was about half of that for the wild-type enzyme. (1)H-NMR analysis of R114H and R114A indicated that the structural changes induced by the mutations were not restricted to the region surrounding Arg114, but rather extended to the aromatic side chains of Phe34 and Trp123, of which the signals are connected with each other through nuclear Overhauser effect (NOE) in the wild-type. We speculate that such a conformational change causes differences in substrate and acceptor binding at subsites E and F, lowering the efficiency of glycosyl transfer reaction of lysozyme.  相似文献   

15.
The contributions of the -3 subsite and a putative +3 subsite to substrate positioning in ChiA from Serratia marcescens have been investigated by comparing how ChiA and its -3 subsite mutant W167A interact with soluble substrates. The data show that Trp - GlcNAc stacking in the -3 subsite rigidifies the protein backbone supporting the formation of the intermolecular interaction network that is necessary for the recognition and positioning of the N-acetyl groups before the -1 subsite. The +3 subsite exhibits considerable substrate affinity that may promote endo-activity in ChiA and/or assist in expelling dimeric products from the +1 and +2 subsites during processive hydrolysis.  相似文献   

16.
We have investigated the mechanism of the interaction of Streptomyces sp. N174 chitosanase with glucosamine hexasaccharide [(GlcN)(6)] by site-directed mutagenesis, thermal unfolding, and (GlcN)(6) digestion experiments, followed by theoretical calculations. From the energy-minimized model of the chitosanase-(GlcN)(6) complex structure (Marcotte et al., 1996), Asp57, which is present in all known chitosanases, was proposed to be one of the amino acid residues that interacts with the oligosaccharide substrate. The chitosanase gene was mutated at Asp57 to Asn (D57N) and Ala (D57A), and the relative activities of the mutated chitosanases were found to be 72 and 0.5% of that of the wild type, respectively. The increase in the transition temperature of thermal unfolding (T(m)), usually observed upon the addition of (GlcN)(n) to chitosanase mutants unaffected in terms of substrate binding, was considerably suppressed in the D57A mutant. These data suggest that Asp57 is important for substrate binding. The experimental time-courses of [(GlcN)(6)] degradation were analyzed by a theoretical model in order to obtain the binding free energy values of the individual subsites of the chitosanases. A (-3, -2, -1, +1, +2, +3) subsite model agreed best with the experimental data. This analysis also indicated that the mutation of Asp57 affects substrate affinity at subsite (-2), suggesting that Asp57 most likely participates in the substrate binding at this subsite.  相似文献   

17.
Kinetic analyses of the hydrolysis reactions of N-acetylated and N-deacetylated derivatives of 4-methylumbelliferyl chitobioside [(GlcNAc)(2)-UMB (1), GlcN-GlcNAc-UMB (2), GlcNAc-GlcN-UMB (3), and (GlcN)(2)-UMB (4)] by ChiA and ChiB from Serratia marcescens were performed. Both enzymes released UMB from all compounds apart from 4. The S-v curves of the hydrolyses of 1 by ChiA and ChiB both exhibited atypical kinetic patterns, and the shapes of the two S-v curves were different from one another. However, both curve shapes were explained by assuming some of the enzyme present formed complexes with multiple molecules of the substrate. Conversely, the S-v curves generated in the cleavage of 2 and 3 by ChiA exhibited typical Michaelis-Menten profiles. Both enzymes hydrolysed 2 with an approximately 14-fold higher K(m) value relative to 1, indicating that the N-acetyl group was recognised at the -2 subsite. The k(cat) value obtained with ChiA was identical to the k(cat) value observed for 1. However, the k(cat) value for ChiB was one-fourth that of 1, suggesting that the removal of the N-acetyl group caused an increase in the formation of a non-productive ES-complex. ChiA and ChiB hydrolysed 3 with 5- and 20-fold greater K(m) values relative to 1, respectively, and 60- and 30-fold smaller k(cat) values relative to 1, respectively. The reaction mechanism of family 18 chitinases is discussed based upon the results obtained from the hydrolysis of these compounds.  相似文献   

18.
Arginine 115 in the subsite F of human lysozyme (peptidoglycan N-acetylmuramoylhydrolase, EC 3.2.1.17) was replaced with lysine, histidine, glutamine or glutamine acid by site-directed mutagenesis. The conversions which conserve positive charge, Arg115 to Lys or His (at acidic pH), have little affected on either the kinetic parameters for Micrococcus lysodeikticus cells or the activity against glycol chitin, nor on the cleavage patterns of hexa(N-acetylglucosamine) [(GlcNAc)6] and penta(N-acetylglucosamine) [(GlcNAc)5]. On the other hand, the conversions which cause loss of the positive charge, Arg115 to His (neutral and alkaline pH), Gln or Glu, not only reduced the activity against glycol chitin but also changed the cleavage patterns for (GlcNAc)6 and (GlcNAc)5. These results suggest that Arg115 is structurally required not for the specific hydrogen bonding interaction with a sugar residue but for the positively charged character in the construction of subsite F in human lysozyme.  相似文献   

19.
We have studied the kinetics of "unisite" ATP hydrolysis and synthesis in seven mutant Escherichia coli F1-ATPase enzymes. The seven mutations are distributed over a 105-residue segment of the catalytic nucleotide-binding domain in beta-subunit and are: G142S, K155Q, K155E, E181Q, E192Q, M209I, and R246C. We report forward and reverse rate constants and equilibrium constants in all seven mutant enzymes for the four steps of unisite kinetics, namely (i) ATP binding/release, (ii) ATP hydrolysis/synthesis, (iii) Pi release/binding, and (iv) ADP release/binding. The seven mutant enzymes displayed a wide range of deviations from normal in both rate and equilibrium constants, with no discernible common pattern. Notably, steep reductions in Kd ATP were seen in some cases, the value of Kd Pi was high, and K2 (ATP hydrolysis/synthesis) was relatively unaffected. Significantly, when the data from the seven mutations were combined with previous data from two other E. coli F1-beta-subunit mutations (D242N, D242V), normal E. coli F1, soluble and membranous mitochondrial F1, it was found that linear free energy relationships obtained for both ATP binding/release (log k+1 versus log K1) and ADP binding/release (log k-4 versus log K-4). Two conclusions follow. 1) The seven mutations studied here cause subtle changes in interactions between the catalytic nucleotide-binding domain and substrate ATP or product ADP. 2) The mitochondrial, normal E. coli, and nine total beta-subunit mutant enzymes represent a continuum in which subtle structural differences in the catalytic site resulted in changes in binding energy; therefore insights into the nature of energy coupling during ATP hydrolysis and synthesis by F1-ATPase may be ascertained by detailed studies of this group of enzymes.  相似文献   

20.
The kinetic behavior of chitinase A1 from Bacillus circulans WL-12 was investigated using the novel fluorogenic substrates, N-deacetylated 4-methylumbelliferyl chitobiosides [GlcN-GlcNAc-UMB (2), GlcNAc-GlcN-UMB (3), and (GlcN)(2)-UMB (4)], and the results were compared with those obtained using 4-methylumbelliferyl N, N'-diacetylchitobiose [(GlcNAc)(2)-UMB (1)] as the substrate. The chitinase did not release the UMB moiety from compound 4, but successfully released UMB from the other substrates. k(cat)/K(m) values determined from the releasing rate of the UMB moiety were: 145.3 for 1, 8.3 for 2, and 0.1 s(-1) M(-1) for 3. The lack of an N-acetyl group at subsite (-1) reduced the activity to a level 0.1% of that obtained with compound 1, while the absence of the N-acetyl group at subsite (-2) reduced the relative activity to 5.7%. These observations strongly support the theory that chitinase A1 catalysis occurs via a 'substrate-assisted' mechanism. Using these novel fluorogenic substrates, we were able to quantitatively evaluate the recognition specificity of subsite (-2) toward the N-acetyl group of the substrate sugar residue. The (-2) subsite of chitinase A1 was found to specifically recognize an N-acetylated sugar residue, but this specificity was not as strict as that found in subsite (-1).  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