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 18 毫秒
1.
Phenylalanine ammonia-lyase (EC 4.3.1.5) from parsley is posttranslationally modified by dehydrating its Ser-202 to the catalytically essential dehydroalanine prosthetic group. The codon of Ser-202 was changed to those of alanine and threonine by site-directed mutagenesis. These mutants and the recombinant wild-type enzyme, after treatment with sodium borohydride, were virtually inactive with L-phenylalanine as substrate but catalyzed the deamination of L-4-nitrophenylalanine, which is also a substrate for the wild-type enzyme. Although the mutants reacted about 20 times slower with L-4-nitrophenylalanine than the wild-type enzyme, their Vmax for L-4-nitrophenylalanine was two orders of magnitude higher than for L-phenylalanine. In contrast to L-tyrosine, which was a poor substrate, DL-3-hydroxyphenylalanine (DL-m-tyrosine) was converted by phenylalanine ammonia-lyase at a rate comparable to that of L-phenylalanine. These results suggest a mechanism in which the crucial step is an electrophilic attack of the prosthetic group at position 2 or 6 of the phenyl group. In the resulting carbenium ion, the beta-HSi atom is activated in a similar way as it is in the nitro analogue. Subsequent elimination of ammonia, concomitant with restoration of both the aromatic ring and the prosthetic group, completes the catalytic cycle.  相似文献   

2.
Phenylalanine dehydrogenase from Thermoactinomyces intermedius and leucine dehydrogenase from Bacillus stearothermophilus show a 59% sequence similarity in their substrate-binding domains, although their substrate specificities are different. We prepared a phenylalanine dehydrogenase mutant enzyme whose inherent hexapeptide segment (124F-V-H-A-A-129R) in the substrate-binding domain was replaced by the corresponding part of leucine dehydrogenase (M-D-I-I-Y-Q) in order to investigate the mechanism of substrate recognition by phenylalanine dehydrogenase. The catalytic efficiencies (kcat/Km) of the mutant enzyme with aliphatic amino acids and aliphatic keto acids as substrates were 0.5 to 2% of those of the wild-type enzyme. In contrast, the efficiencies for L-phenylalanine and phenylpyruvate decreased to 0.008 and 0.035% of those of the wild-type enzyme, respectively. These results suggest that the hexapeptide segment plays an important role in the substrate recognition by phenylalanine dehydrogenase.  相似文献   

3.
Aromatase is a cytochrome P450 enzyme complex that catalyzes the conversion of androst-4-ene-3,17-dione to estrone through three sequential oxygenations of the 19-methyl group. Androst-4-ene-3,6,17-trione (1) is a suicide substrate of aromatase. The inactivation mechanism for steroid 1 has been studied to show that the inactivation reaction proceeds through the 19-oxo intermediate 3. To further clarify the mechanism, 4 beta, 5 beta-epoxyandrosta-3,6,17,19-tetraone (6) was synthesized as a candidate for a reactive electrophile involved in irreversible binding to the active site of aromatase, upon treatment of compound 3 with hydrogen peroxide in the presence of NaHCO3. The epoxide 6 inhibited human placental aromatase in a competitive manner (Ki = 30 microM); moreover, it inactivated the enzyme in an active-site-directed manner in the absence of NADPH (K1 = 88 microM, kinact = 0.071 min-1). NADPH and BSA both stimulated the inactivation rate without a significant change of the K1 in either case (kinact: 0.133 or 0.091 min-1, in the presence of NADPH or BSA, respectively). The substrate androst-4-ene-3,17-dione protected the inactivation, but a nucleophile, L-cysteine, did not. When both the epoxide 6 and its 19-methyl analog 4 were subjected separately to reaction with N-acetyl-L-cysteine in the presence of NaHCO3, the 19-oxo steroid 6 disappeared from the reaction mixture more rapidly (T1/2 = 40 sec) than the 19-methyl analog 4 (T1/2 = 3.0 min). The results clearly indicate that the 4 beta, 5 beta-epoxy-19-oxo compound 6, which is possibly produced from 19-oxo-4-ene steroid 3 through the 19-hydroxy-19-hydroperoxide intermediate, is a reactive electrophile that irreversibly binds to the active site of aromatase.  相似文献   

4.
Novel information on the structure and function of chitosanase, which hydrolyzes the beta-1,4-glycosidic linkage of chitosan, has accumulated in recent years. The cloning of the chitosanase gene from Streptomyces sp. strain N174 and the establishment of an efficient expression system using Streptomyces lividans TK24 have contributed to these advances. Amino acid sequence comparisons of the chitosanases that have been sequenced to date revealed a significant homology in the N-terminal module. From energy minimization based on the X-ray crystal structure of Streptomyces sp. strain N174 chitosanase, the substrate binding cleft of this enzyme was estimated to be composed of six monosaccharide binding subsites. The hydrolytic reaction takes place at the center of the binding cleft with an inverting mechanism. Site-directed mutagenesis of the carboxylic amino acid residues that are conserved revealed that Glu-22 and Asp-40 are the catalytic residues. The tryptophan residues in the chitosanase do not participate directly in the substrate binding but stabilize the protein structure by interacting with hydrophobic and carboxylic side chains of the other amino acid residues. Structural and functional similarities were found between chitosanase, barley chitinase, bacteriophage T4 lysozyme, and goose egg white lysozyme, even though these proteins share no sequence similarities. This information can be helpful for the design of new chitinolytic enzymes that can be applied to carbohydrate engineering, biological control of phytopathogens, and other fields including chitinous polysaccharide degradation.  相似文献   

5.
Gln34, Gln224, Leu228, and Ser240 are conserved residues in the vicinity of bound IMP in the crystal structure of Escherichia coli adenylosuccinate synthetase. Directed mutations were carried out, and wild-type and mutant enzymes were purified to homogeneity. Circular dichroism spectroscopy indicated no difference in secondary structure between the mutants and the wild-type enzyme in the absence of substrates. Mutants L228A and S240A exhibited modest changes in their initial rate kinetics relative to the wild-type enzyme, suggesting that neither Leu228 nor Ser240 play essential roles in substrate binding or catalysis. The mutants Q224M and Q224E exhibited no significant change in KmGTP and KmASP and modest changes in KmIMP relative to the wild-type enzyme. However, kcat decreased 13-fold for the Q224M mutant and 10(4)-fold for the Q224E mutant relative to the wild-type enzyme. Furthermore, the Q224E mutant showed an optimum pH at 6.2, which is 1.5 pH units lower than that of the wild-type enzyme. Tryptophan emission fluorescence spectra of Q224M, Q224E, and wild-type proteins under denaturing conditions indicate comparable stabilities. Mutant Q34E exhibits a 60-fold decrease in kcat compared with that of the wild-type enzyme, which is attributed to the disruption of the Gln34 to Gln224 hydrogen bond observed in crystal structures. Presented here is a mechanism for the synthetase, whereby Gln224 works in concert with Asp13 to stabilize the 6-oxyanion of IMP.  相似文献   

6.
Crystal structures of L-2-haloacid dehalogenase from Pseudomonas sp. YL complexed with monochloroacetate, L-2-chlorobutyrate, L-2-chloro-3-methylbutyrate, or L-2-chloro-4-methylvalerate were determined at 1.83-, 2.0-, 2.2-, and 2.2-A resolutions, respectively, using the complex crystals prepared with the S175A mutant, which are isomorphous with those of the wild-type enzyme. These structures exhibit unique structural features that correspond to those of the reaction intermediates. In each case, the nucleophile Asp-10 is esterified with the dechlorinated moiety of the substrate. The substrate moieties in all but the monochloroacetate intermediate have a D-configuration at the C2 atom. The overall polypeptide fold of each of the intermediates is similar to that of the wild-type enzyme. However, it is clear that the Asp-10-Ser-20 region moves to the active site in all of the intermediates, and the Tyr-91-Asp-102 and Leu-117-Arg-135 regions make conformational changes in all but the monochloroacetate intermediates. Ser-118 is located near the carboxyl group of the substrate moiety; this residue probably serves as a binding residue for the substrate carboxyl group. The hydrophobic pocket, which is primarily composed of the Tyr-12, Gln-42, Leu-45, Phe-60, Lys-151, Asn-177, and Trp-179 side chains, exists around the alkyl group of the substrate moiety. This pocket may play an important role in stabilizing the alkyl group of the substrate moiety through hydrophobic interactions, and may also play a role in determining the stereospecificity of the enzyme. Moreover, a water molecule, which is absent in the substrate-free enzyme, is present in the vicinities of the carboxyl carbon of Asp-10 and the side chains of Asp-180, Asn-177, and Ala-175 in each intermediate. This water molecule may hydrolyze the ester intermediate and its substrate. These findings crystallographically demonstrate that the enzyme reaction proceeds through the formation of an ester intermediate with the enzyme's nucleophile Asp-10.  相似文献   

7.
Degradation of a protein via the ubiquitin system involves two discrete steps, conjugation of ubiquitin to the substrate and degradation of the adduct. Conjugation follows a three-step mechanism. First, ubiquitin is activated by the ubiquitin-activating enzyme, E1. Following activation, one of several E2 enzymes (ubiquitin-carrier proteins or ubiquitin-conjugating enzymes, UBCs) transfers ubiquitin from E1 to the protein substrate that is bound to one of several ubiquitin-protein ligases, E3s. These enzymes catalyze the last step in the process, covalent attachment of ubiquitin to the protein substrate. The binding of the substrate to E3 is specific and implies that E3s play a major role in recognition and selection of proteins for conjugation and subsequent degradation. So far, only a few ligases have been identified, and it is clear that many more have not been discovered yet. Here, we describe a novel ligase that is involved in the conjugation and degradation of non "N-end rule" protein substrates such as actin, troponin T, and MyoD. This substrate specificity suggests that the enzyme may be involved in degradation of muscle proteins. The ligase acts in concert with E2-F1, a previously described non N-end rule UBC. Interestingly, it is also involved in targeting lysozyme, a bona fide N-end substrate that is recognized by E3 alpha and E2-14 kDa. The novel ligase recognizes lysozyme via a signal(s) that is distinct from the N-terminal residue of the protein. Thus, it appears that certain proteins can be targeted via multiple recognition motifs and distinct pairs of conjugating enzymes. We have purified the ligase approximately 200-fold and demonstrated that it is different from other known E3s, including E3 alpha/UBR1, E3 beta, and E6-AP. The native enzyme has an apparent molecular mass of approximately 550 kDa and appears to be a homodimer. Because of its unusual size, we designated this novel ligase E3L (large). E3L contains an -SH group that is essential for its activity. Like several recently described E3 enzymes, including E6-AP and the ligase involved in the processing of p105, the NF-kappa B precursor, the novel ligase is found in mammalian tissues but not in wheat germ.  相似文献   

8.
采用微弧氧化技术,在TC4钛合金表面制备高硬度氧化陶瓷层(MAO),对比研究了TC4钛合金基体与微弧氧化陶瓷层在2种不同位移幅值下的微动磨损行为.结果表明:位移幅值由80μm增大到150μm时,TC4钛合金基体微动损伤机制由粘着磨损和磨粒磨损转变为疲劳磨损和氧化磨损,而微弧氧化陶瓷层的损伤机制始终以氧化磨损为主;位移幅...  相似文献   

9.
L-2-Haloacid dehalogenase (EC 3.8.1.2) catalyzes the hydrolytic dehalogenation of L-2-haloacids to produce the corresponding D-2-hydroxy acids. We have analyzed the reaction mechanism of the enzyme from Pseudomonas sp. YL and found that Asp10 is the active site nucleophile. When the multiple turnover enzyme reaction was carried out in H2(18)O with L-2-chloropropionate as a substrate, lactate produced was labeled with 18O. However, when the single turnover enzyme reaction was carried out by use of a large excess of the enzyme, the product was not labeled. This suggests that an oxygen atom of the solvent water is first incorporated into the enzyme and then transferred to the product. After the multiple turnover reaction in H2(18)O, the enzyme was digested with lysyl endopeptidase, and the molecular masses of the peptide fragments formed were measured by an ionspray mass spectrometer. Two 18O atoms were shown to be incorporated into a hexapeptide, Gly6-Lys11. Tandem mass spectrometric analysis of this peptide revealed that Asp10 was labeled with two 18O atoms. Our previous site-directed mutagenesis experiment showed that the replacement of Asp10 led to a significant loss in the enzyme activity. These results indicate that Asp10 acts as a nucleophile on the alpha-carbon of the substrate leading to the formation of an ester intermediate, which is hydrolyzed by nucleophilic attack of a water molecule on the carbonyl carbon atom.  相似文献   

10.
We have devised an assay procedure that permits simultaneous monitoring of the four activities of ribonucleotide reductase. Using this assay, we have compared the reduction of all four substrates by the T4 bacteriophage aerobic ribonucleotide reductase within different allosteric environments. Specifically, we compared the relative turnover rates by the enzyme when activated with "in vivo" concentrations of the known allosteric effectors versus activation by ATP alone. Consistent with the known allosteric properties of this enzyme, our results show that ATP does act as a general activator, although the rate of purine nucleotide reduction was approximately 5% of the rate for the pyrimidine nucleotides. However, addition of the allosteric effectors at their estimated physiological concentrations dramatically changed the relative rates of substrate reduction, creating a more "balanced" pool of products. Addition of the substrates at their respective in vivo concentrations further pushed rates of product formation toward a ratio similar to the base composition of the T4 genome. The similarity of the product profile produced under in vivo conditions to the genomic composition of T4 phage is discussed.  相似文献   

11.
A method for evaluating the association and dissociation rate constants of interaction between a lysozyme and its substrate analogue, an immobilized p-aminophenyl-tri-N-acetyl-beta-chitotrioside, by means of surface plasmon resonance has been developed. Site-specific immobilization of p-aminophenyl-tri-N-acetyl-beta-chitotrioside, which is a product of p-nitrophenyl-tri-N-acetyl-beta-chitotrioside, on carboxymethyldextran linked to the surface of the cuvette of the instrument, IAsys, was carried out by catalysis with EDC/NHS. The kinetic parameters of the interaction between hen or human lysozyme and the immobilized substrate analogue indicated that a larger dissociation constant of the human lysozyme-immobilized substrate analogue complex depended on a smaller association rate constant. The kinetic parameters of the interaction between the immobilized substrate analogue and a mutant hen lysozyme, in which Arg14 and His15 are deleted, with higher activity than the wild type hen lysozyme were measured. It was suggested that the higher activity of the mutant lysozyme was due to faster removal of the substrate from the active site cleft and/or the formation of a stabler and better complex as to hydrolysis.  相似文献   

12.
The carbohydrate-binding cleft of Bacillus licheniformis 1,3-1, 4-beta-D-glucan 4-glucanohydrolase is partially covered by the surface loop between residues 51 and 67, which is linked to beta-strand-(87-95) of the minor beta-sheet III of the protein core by a single disulfide bond at Cys61-Cys90. An alanine scanning mutagenesis approach has been applied to analyze the role of loop residues from Asp51 to Arg64 in substrate binding and stability by means of equilibrium urea denaturation, enzyme thermotolerance, and kinetics. The DeltaDeltaGU between oxidized and reduced forms is approximately constant for all mutants, with a contribution of 5.3 +/- 0.2 kcal.mol-1 for the disulfide bridge to protein stability. A good correlation is observed between DeltaGU values by reversible unfolding and enzyme thermotolerance. The N57A mutant, however, is more thermotolerant than the wild-type enzyme, whereas it is slightly less stable to reversible urea denaturation. Mutants with a <2-fold increase in Km correspond to mutations at residues not involved in substrate binding, for which the reduction in catalytic efficiency (kcat/Km) is proportional to the loss of stability relative to the wild-type enzyme. Y53A, N55A, F59A, and W63A, on the other hand, show a pronounced effect on catalytic efficiency, with Km > 2-fold and kcat < 5% of the wild-type values. These mutated residues are directly involved in substrate binding or in hydrophobic packing of the loop. Interestingly, the mutation M58A yields an enzyme that is more active than the wild-type enzyme (7-fold increase in kcat), but it is slightly less stable.  相似文献   

13.
DNA (cytosine-5-)-methyltransferase is essential for viable mammalian development and has a central function in the determination and maintenance of epigenetic methylation patterns. Steady-state and substrate trapping studies were performed to better understand how the enzyme functions. The catalytic efficiency was dependent on substrate DNA length. A 14-fold increase in KmDNA was observed as the length decreased from 5000 to 100 base pairs and kcat decreased by a third. Steady-state analyses were used to identify the order of substrate addition onto the enzyme and the order of product release. Double-reciprocal patterns of velocity versus substrate concentration intersected far from the origin and were nearly parallel. The kinetic mechanism does not appear to change when the DNA substrate is either 6250 or 100 base pairs in length. Isotope trapping studies showed that the initial enzyme-AdoMet complex was not catalytically competent; however, the initial enzyme-poly(dI.dC-dI.dC) complex was observed to be competent for catalysis. Product inhibition studies also support a sequential ordered bi-bi kinetic mechanism in which DNA binds to the enzyme first, followed by S-adenosyl-L-methionine, and then the products S-adenosyl-L-homocysteine and methylated DNA are released. The proposed mechanism is similar to the mechanism proposed for M. HhaI, a bacterial DNA (cytosine-5-)-methyltransferase. Evidence for an enzyme-DNA-DNA ternary complex is also presented.  相似文献   

14.
Comparisons of the structures of glutamate dehydrogenase (GluDH) and leucine dehydrogenase (LeuDH) have suggested that two substitutions, deep within the amino acid binding pockets of these homologous enzymes, from hydrophilic residues to hydrophobic ones are critical components of their differential substrate specificity. When one of these residues, K89, which hydrogen-bonds to the gamma-carboxyl group of the substrate l-glutamate in GluDH, was altered by site-directed mutagenesis to a leucine residue, the mutant enzyme showed increased substrate activity for methionine and norleucine but negligible activity with either glutamate or leucine. In order to understand the molecular basis of this shift in specificity we have determined the crystal structure of the K89L mutant of GluDH from Clostridium symbiosum. Analysis of the structure suggests that further subtle differences in the binding pocket prevent the mutant from using a branched hydrophobic substrate but permit the straight-chain amino acids to be used as substrates. The three-dimensional crystal structure of the GluDH from C. symbiosum has been previously determined in two distinct forms in the presence and absence of its substrate glutamate. A comparison of these two structures has revealed that the enzyme can adopt different conformations by flexing about the cleft between its two domains, providing a motion which is critical for orienting the partners involved in the hydride transfer reaction. It has previously been proposed that this conformational change is triggered by substrate binding. However, analysis of the K89L mutant shows that it adopts an almost identical conformation with that of the wild-type enzyme in the presence of substrate. Comparison of the mutant structure with both the wild-type open and closed forms has enabled us to separate conformational changes associated with substrate binding and domain motion and suggests that the domain closure may well be a property of the wild-type enzyme even in the absence of substrate.  相似文献   

15.
Transgenic mice (T26) bearing the envelope, regulatory, and accessory genes of HIV- I develop renal disease resembling human HIV-associated nephropathy (HIVAN). Effects of vehicle (VEH) and the angiotensin-converting enzyme inhibitor captopril (CAP) were examined in wild-type (WT) or T26 mice treated from 7 to 100 d of age. Mortality was lower in CAP T26 mice (30 mg/kg: 8%; 100 mg/kg: 12%) than VEH T26 mice (52%). The urinary protein/creatinine ratio was increased in VEH T26 mice (19.5+/-7.60) versus WT mice (6.1+/-0.83), but not in low-dose (7.3+/-0.94) or high-dose (8.2+/-1.02) CAP T26 mice. Blood urea nitrogen was higher in VEH T26 mice (52+/-16.2 mg/dl) than VEH WT mice (24+/-0.8). Blood urea nitrogen was also elevated in CAP WT (high dose: 43+/-2.1 mg/dl) and T26 mice (high dose: 42+/-2.4 mg/dl). Glomerular injury was higher in VEH T26 mice (6.8+/-0.58) than VEH WT mice (0.2+/-0.08) or CAP T26 mice (low dose: 1.1+/-0.17; high dose: 0.7+/-0.13). Tubulointerstitial injury was also greater in VEH T26 mice (1.1+/-0.10) than VEH WT mice (0.2+/-0.08) or CAP T26 mice (low dose: 0.4+/-0.10; high dose: 0.3+/-0.10). These data validate recent nonrandomized studies of captopril in HIV-infected patients, and suggest that an angiotensin-converting enzyme substrate is an important mediator in HIVAN. A randomized placebo-controlled trial of captopril in HIVAN may be warranted.  相似文献   

16.
A lysozyme (pI 5.5) was purified to homogeneity from heated acid extracts of Drosophila melanogaster larvae, using gel filtration in a Superose column and ion-exchange chromatography in a Mono Q column. The final yield was 67%. The purified lysozyme with Mr 13,700 (determined by SDS-polyacrylamide gel electrophoresis) decreases in activity and has its pH optimum displaced towards acidic values and Km increases as the ionic strength of the medium becomes higher. The lysozyme is resistant to a cathepsin D-like proteinase present in cyclorrhaphous Diptera and displays a chitinase activity which is 11-fold higher than that of chicken lysozyme. Microsequencing of an internal peptide of the purified lysozyme showed that this enzyme is the product of the previously sequenced Lys D gene. The results suggest that the product of the Lys P gene has pI 7.2, a pH optimum around 5 and is not a true digestive enzyme. The most remarkable sequence convergence of D. melanogaster lysozyme D and lysozymes from vertebrate foregut fermenters are serine 104 and a decrease in the number of basic amino acids, suggesting that these features are necessary for digestive function in an acid environment. Adaptive residues putatively conferring stability in an acid proteolytic environment differ between insects and vertebrates, probably because they depend on the overall three-dimensional structure of the lysozymes. A maximum likelihood phylogeny and inferences from insect lysozyme sequences showed that the recruitment of lysozymes as digestive enzymes is an ancestral condition of the flies (Diptera: Cyclorrhapha).  相似文献   

17.
Glutamine synthetase (GS) catalyzes the ATP-dependent condensation of ammonia and glutamate to yield glutamine, ADP, and inorganic phosphate in the presence of divalent cations. Bacterial GS is an enzyme of 12 identical subunits, arranged in two rings of 6, with the active site between each pair of subunits in a ring. In earlier work, we have reported the locations within the funnel-shaped active site of the substrates glutamate and ATP and of the two divalent cations, but the site for ammonia (or ammonium) has remained elusive. Here we report the discovery by X-ray crystallography of a binding site on GS for monovalent cations, Tl+ and Cs+, which is probably the binding site for the substrate ammonium ion. Fourier difference maps show the following. (1) Tl+ and Cs+ bind at essentially the same site, with ligands being Glu 212, Tyr 179, Asp 50', Ser 53' of the adjacent subunit, and the substrate glutamate. From its position adjacent to the substrate glutamate and the cofactor ADP, we propose that this monovalent cation site is the substrate ammonium ion binding site. This proposal is supported by enzyme kinetics. Our kinetic measurements show that Tl+, Cs+, and NH4+ are competitive inhibitors to NH2OH in the gamma-glutamyl transfer reaction. (2) GS is a trimetallic enzyme containing two divalent cation sites (n1, n2) and one monovalent cation site per subunit. These three closely spaced ions are all at the active site: the distance between n1 and n2 is 6 A, between n1 and Tl+ is 4 A, and between n2 and Tl+ is 7 A. Glu 212 and the substrate glutamate are bridging ligands for the n1 ion and Tl+. (3) The presence of a monovalent cation in this site may enhance the structural stability of GS, because of its effect of balancing the negative charges of the substrate glutamate and its ligands and because of strengthening the "side-to-side" intersubunit interaction through the cation-protein bonding. (4) The presence of the cofactor ADP increases the Tl+ binding to GS because ADP binding induces movement of Asp 50' toward this monovalent cation site, essentially forming the site. This observation supports a two-step mechanism with ordered substrate binding: ATP first binds to GS, then Glu binds and attacks ATP to form gamma-glutamyl phosphate and ADP, which complete the ammonium binding site. The third substrate, an ammonium ion, then binds to GS, and then loses a proton to form the more active species ammonia, which attacks the gamma-glutamyl phosphate to yield Gln. (5) Because the products (Glu or Gln) of the reactions catalyzed by GS are determined by the molecule (water or ammonium) attacking the intermediate gamma-glutamyl phosphate, this negatively charged ammonium binding pocket has been designed naturally for high affinity of ammonium to GS, permitting glutamine synthesis to proceed in aqueous solution.  相似文献   

18.
A genomic library of Pseudomonas fluorescens subsp. cellulosa DNA was screened for galactanase-positive recombinants. The nine galactanase positive phage isolated contained the same galactanase gene designated galA. The deduced primary structure of the enzyme (galactanase A; GalA) encoded by galA had a Mr of 42 130 and exhibited significant sequence identity with a galactanase from Aspergillus aculeatus, placing GalA in glycosyl hydrolase family 53. The enzyme displayed properties typical of an endo-beta1, 4-galactanase and exhibited no activity against the other plant structural polysaccharides evaluated. Analysis of the stereochemical course of 2,4-dinitrophenyl-beta-galactobioside (2,4-DNPG2) hydrolysis by GalA indicated that the galactanase catalyzes the hydrolysis of glycosidic bonds by a double displacement general acid-base mechanism. Hydrophobic cluster analysis (HCA) suggested that family 53 enzymes are related to the GH-A clan of glycosyl hydrolases, which have an (alpha/beta)8 barrel structure. HCA also predicted that E161 and E270 were the acid-base and nucleophilic residues, respectively. Mutants of GalA in which E161 and E270 had been replaced with alanine residues were essentially inactive against galactan. Against 2,4-DNPG2, E161A exhibited a much lower Km and kcat than native GalA, while E270A was inactive against the substrate. Analysis of the pre-steady-state kinetics of 2,4-DNPG2 hydrolysis by E161A showed that there was an initial rapid release of 2,4-dinitrophenol (2,4-DNP), which then decayed to a slow steady-state rate of product formation. No pre-steady-state burst of 2,4-DNP release was observed with the wild-type enzyme. These data are consistent with the HCA prediction that E161 and E270 are the acid-base and nucleophilic catalytic residues of GalA, respectively.  相似文献   

19.
Mutant adenylosuccinate lyases of Bacillus subtilis were prepared by site-directed mutagenesis with replacements for His141, previously identified by affinity labeling as being in the active site [Lee, T. T., Worby, C., Dixon, J. E., and Colman, R. F. (1997) J. Biol. Chem. 272, 458-465]. Four substitutions (A, L, E, Q) yield mutant enzyme with no detectable catalytic activity, while the H141R mutant is about 10(-)5 as active as the wild-type enzyme. Kinetic studies show, for the H141R enzyme, a Km that is only 3 times that of the wild-type enzyme. Minimal activity was also observed for mutant enzymes with replacements for His68 [Lee, T. T., Worby, C., Bao, Z. -Q., Dixon, J. E., and Colman, R. F. (1998) Biochemistry 37, 8481-8489]. Measurement of the reversible binding of radioactive adenylosuccinate by inactive mutant enzymes with substitutions at either position 68 or 141 shows that their affinities for substrate are decreased by only 10-40-fold. These results suggest that His141, like His68, plays an important role in catalysis, but not in substrate binding. Evidence is consistent with the hypothesis that His141 and His68 function, respectively, as the catalytic base and acid. Circular dichroism spectroscopy and gel filtration chromatography conducted on wild-type and all His141 and His68 mutants reveal that none of the mutant enzymes exhibits major structural changes and that all the enzymes are tetramers. Mixing inactive His141 with inactive His68 mutant enzymes leads to striking increases in catalytic activity. This complementation of mutant enzymes indicates that His141 and His68 come from different subunits to form the active site. A tetrameric structure of adenylosuccinate lyase was constructed by homology modeling based on the known structures in the fumarase superfamily, including argininosuccinate lyase, delta-crystallin, fumarase, and aspartase. The model suggests that each active site is constituted by residues from three subunits, and that His141 and His68 come from two different subunits.  相似文献   

20.
The lysine residue binding with the cofactor pyridoxal 5'-phosphate (PLP) plays an important role in catalysis, such as in the transaldimination and abstraction of alpha-hydrogen from a substrate amino acid in PLP-dependent enzymes. We studied the role of Lys39 of alanine racemase (EC 5.1.1.1) from Bacillus stearothermophilus, the PLP-binding residue of the enzyme, by replacing it site-specifically with alanine and characterizing the resultant K39A mutant enzyme. The mutant enzyme turned out to be inherently inactive, but gained an activity as high as about 0.1% of that of the wild-type enzyme upon addition of 0.2 M methylamine. The amine-assisted activity of the mutant enzyme depended on the pKa values and molecular volumes of the alkylamines used. A strong kinetic isotope effect was observed when alpha-deuterated D-alanine was used as a substrate in the methylamine-assisted reaction, but little effect was observed using its antipode. In marked contrast, only L-enantiomer of alanine showed a solvent isotope effect in deuterium oxide in the methylamine-assisted reaction. These results suggest that methylamine serves as a base not only to abstract the alpha-hydrogen from D-alanine but also to transfer a proton from water to the alpha-position of the deprotonated (achiral) intermediate to form D-alanine. Therefore, the exogenous amine can be regarded as a functional group fully representing Lys39 of the wild-type enzyme. Lys39 of the wild-type enzyme probably acts as the base catalyst specific to the D-enantiomer of alanine. Another residue specific to the L-enantiomer in the wild-type enzyme is kept intact in the K39A mutant.  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