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 0 毫秒
1.
Aldose reductase (AR), a member of the aldo-keto reductase superfamily, has been implicated in the etiology of secondary diabetic complications. However, the physiological functions of AR under euglycemic conditions remain unclear. We have recently demonstrated that, in intact heart, AR catalyzes the reduction of the glutathione conjugate of the lipid peroxidation product 4-hydroxy-trans-2-nonenal (Srivastava, S., Chandra, A., Wang, L., Seifert, W. E., Jr., DaGue, B. B., Ansari, N. H., Srivastava, S. K., and Bhatnagar, A. (1998) J. Biol. Chem. 273, 10893-10900), consistent with a possible role of AR in the metabolism of glutathione conjugates of aldehydes. Herein, we present several lines of evidence suggesting that the active site of AR forms a specific glutathione-binding domain. The catalytic efficiency of AR in the reduction of the glutathione conjugates of acrolein, trans-2-hexenal, trans-2-nonenal, and trans,trans-2,4-decadienal was 4-1000-fold higher than for the corresponding free alkanal. Alterations in the structure of glutathione diminished the catalytic efficiency in the reduction of the acrolein adduct, consistent with the presence of specific interactions between the amino acid residues of glutathione and the AR active site. In addition, non-aldehydic conjugates of glutathione or glutathione analogs displayed active-site inhibition. Molecular dynamics calculations suggest that the conjugate adopts a specific low energy configuration at the active site, indicating selective binding. These observations support an important role of AR in the metabolism of glutathione conjugates of endogenous and xenobiotic aldehydes and demonstrate, for the first time, efficient binding of glutathione conjugates to an aldo-keto reductase.  相似文献   

2.
The human 8-oxoguanine DNA glycosylase (hOGG1) protein is responsible for initiating base excision DNA repair of the endogenous mutagen 8-oxoguanine. Like nearly all DNA glycosylases, hOGG1 extrudes its substrate from the DNA helix and inserts it into an extrahelical enzyme active site pocket lined with residues that participate in lesion recognition and catalysis. Structural analysis has been performed on mutant versions of hOGG1 having changes in catalytic residues but not on variants having altered 7,8-dihydro-8-oxoguanine (oxoG) contact residues. Here we report high resolution structural analysis of such recognition variants. We found that Ala substitution at residues that contact the phosphate 5' to the lesion (H270A mutation) and its Watson-Crick face (Q315A mutation) simply removed key functionality from the contact interface but otherwise had no effect on structure. Ala substitution at the only residue making an oxoG-specific contact (G42A mutation) introduced torsional stress into the DNA contact surface of hOGG1, but this was overcome by local interactions within the folded protein, indicating that this oxoG recognition motif is "hardwired." Introduction of a side chain intended to sterically obstruct the active site pocket (Q315F mutation) led to two different structures, one of which (Q315F(*149)) has the oxoG lesion in an exosite flanking the active site and the other of which (Q315F(*292)) has the oxoG inserted nearly completely into the lesion recognition pocket. The latter structure offers a view of the latest stage in the base extrusion pathway yet observed, and its lack of catalytic activity demonstrates that the transition state for displacement of the lesion base is geometrically demanding.  相似文献   

3.
5,10-Methylenetetrahydrofolate dehydrogenase (MTD) catalyzes the reversible oxidation of 5,10-methylenetetrahydrofolate to 5,10-methenyltetrahydrofolate. This reaction is critical for the supply of one-carbon units at the required oxidation states for the synthesis of purines and dTMP. For most MTDs, dehydrogenase activity is co-located with a methenyl-THF cyclohydrolase activity as part of bifunctional or trifunctional enzyme. The yeast Saccharomyces cerevisiae contains a monofunctional NAD(+)-dependent 5,10-methylenetetrahydrofolate dehydrogenase (yMTD). Kinetic, crystallographic, and mutagenesis studies were conducted to identify critical residues in order to gain further insight into the reaction mechanism of this enzyme and its apparent lack of cyclohydrolase activity. Hydride transfer was found to be rate-limiting for the oxidation of methylenetetrahydrofolate by kinetic isotope experiments (V(H)/V(D) = 3.3), and the facial selectivity of the hydride transfer to NAD(+) was determined to be Pro-R (A-specific). Model building based on the previously solved structure of yMTD with bound NAD cofactor suggested a possible role for three conserved amino acids in substrate binding or catalysis: Glu121, Cys150, and Thr151. Steady-state kinetic measurements of mutant enzymes demonstrated that Glu121 and Cys150 were essential for dehydrogenase activity, whereas Thr151 allowed some substitution. Our results are consistent with a key role for Glu121 in correctly binding the folate substrate; however, the exact role of C150 is unclear. Single mutants Thr57Lys and Tyr98Gln and double mutant T57K/Y98Q were prepared to test the hypothesis that the lack of cyclohydrolase activity in yMTD was due to the substitution of a conserved Lys/Gln pair found in bifunctional MTDs. Each mutant retained dehydrogenase activity, but no cyclohydrolase activity was detected.  相似文献   

4.
The present study describes the cloning, isolation, and thorough biochemical characterization of UreE from Bacillus pasteurii, a novel protein putatively involved in the transport of Ni in the urease assembly process. A DNA fragment of the B. pasteurii urease operon, containing all four accessory genes (ureE, ureF, ureG, and ureD) required for the incorporation of Ni ions into the active site of urease, was cloned, sequenced, and analyzed. B. pasteurii ureE was cloned, and the UreE protein (BpUreE) was over-expressed and purified to homogeneity. The identity of the recombinant protein was determined by N- and C-terminal sequencing and by mass spectrometry. BpUreE has a chain length of 147 amino acids, and features a p I value of 4.7. As isolated, BpUreE contains one Zn(II) ion per dimer, while no Ni(II) is present, as shown by mass spectrometry and atomic absorption spectroscopy. BpUreE behaves as a dimer independently of the presence of Zn(II), as shown by gel filtration and mass spectrometry. Paramagnetic NMR spectroscopy on concentrated (2 mM) UreE solutions reveals a one Ni atom per tetramer stoichiometry, with the Ni(II) ion bound to histidines in an octahedral coordination environment. BpUreE has a high sequence similarity with UreE proteins isolated from different biological sources, while no sequence homology is observed with proteins belonging to different classes. In particular, BpUreE is most similar to UreE from Bacillus halodurans (55% identity). A multiple sequence alignment reveals the presence of four strictly conserved residues (Leu55, Gly97, Asn98, His100; BpUreE numbering), in addition to position 115, conservatively occupied by an Asp or a Glu residue. Several secondary structure elements, including a betaalphabetabetaalphabeta "ferredoxin-like" motif, are highly conserved throughout the UreE sequences.  相似文献   

5.
A L Fink  K M Behner  A K Tan 《Biochemistry》1987,26(14):4248-4258
The reversible inhibition of beta-lactamase I from Bacillus cereus by cloxacillin, methicillin, and nafcillin has been systematically investigated. For these substrates the enzymatic reaction involves partitioning of the substrate between turnover and inhibition. Typically, concentrations of several hundred millimolar are necessary for complete inactivation. The completely inactivated enzyme could be formed by incubation at temperatures above 20 degrees C, where inhibition competes more effectively with turnover, and then stabilized by dropping the temperature to 0 degrees C or lower. The inactivated enzyme was rapidly separated from unreacted substrate and product at low temperature by centrifugal gel filtration or ion exchange and examined by far-UV circular dichroism for evidence of a conformational change. At pH 7 the inactivated enzyme had a secondary structure essentially identical with that of the native enzyme. The fluorescence emission spectrum of the inactivated enzyme (at pH 7) was also identical with that of the native enzyme. However, the inactivated enzyme was found to be considerably more sensitive to thermal denaturation, to acid-induced conformational isomerization, and to trypsinolysis than the native enzyme. We conclude from the circular dichroism results that the structure of the reversibly inactivated enzyme cannot be significantly different from that of the native enzyme. Therefore, previous findings that have been interpreted as indicating a major conformational change must be reevaluated. From examination of the low-resolution crystallographic structure of the enzyme we propose that the most likely cause of the inactivation is an alternate conformational state of the acyl-enzyme intermediate involving movement of one or more of the alpha-helices forming part of the active site. Such a structural effect could leave the secondary structure unchanged but have significant effects on the tertiary structure, catalysis, mobility, and susceptibility to trypsin and denaturation. We propose that the underlying physical reason why certain beta-lactam substrates bring about this "substrate-induced deactivation", or suicide inactivation, of the enzyme is due to the presence of the alternative acyl-enzyme conformation of similar free energy to the productive one, in which one (or more) essential catalytic group is no longer optimally oriented for catalyzing deacylation. Thus for substrates with a slow rate of deacylation (less than or equal to 100 s-1) the conformational transition can compete effectively on the time scale of the turnover reaction.  相似文献   

6.
Pyridoxal kinase catalyzes the phosphorylation of pyridoxal (PL) to pyridoxal 5′-phosphate (PLP). A D235A variant shows 7-fold and 15-fold decreases in substrate affinity and activity, respectively. A D235N variant shows ∼2-fold decrease in both PL affinity and activity. The crystal structure of D235A (2.5 Å) shows bound ATP, PL and PLP, while D235N (2.3 Å) shows bound ATP and sulfate. These results document the role of Asp235 in PL kinase activity. The observation that the active site of PL kinase can accommodate both ATP and PLP suggests that formation of a ternary Enz·PLP·ATP complex could occur in the wild-type enzyme, consistent with severe MgATP substrate inhibition of PL kinase in the presence of PLP.  相似文献   

7.
Yue QK  Kass IJ  Sampson NS  Vrielink A 《Biochemistry》1999,38(14):4277-4286
Cholesterol oxidase is a monomeric flavoenzyme which catalyzes the oxidation and isomerization of cholesterol to cholest-4-en-3-one. The enzyme interacts with lipid bilayers in order to bind its steroid substrate. The X-ray structure of the enzyme from Brevibacterium sterolicum revealed two loops, comprising residues 78-87 and residues 433-436, which act as a lid over the active site and facilitate binding of the substrate [Vrielink et al. (1991) J. Mol. Biol. 219, 533-554; Li et al. (1993) Biochemistry 32, 11507-11515]. It was postulated that these loops must open, forming a hydrophobic channel between the membrane and the active site of the protein and thus sequestering the cholesterol substrate from the aqueous environment. Here we describe the three-dimensional structure of the homologous enzyme from Streptomyces refined to 1.5 A resolution. Structural comparisons to the enzyme from B. sterolicum reveal significant conformational differences in these loop regions; in particular, a region of the loop comprising residues 78-87 adopts a small amphipathic helical turn with hydrophobic residues directed toward the active site cavity and hydrophilic residues directed toward the external surface of the molecule. It seems reasonable that this increased rigidity reduces the entropy loss that occurs upon binding substrate. Consequently, the Streptomyces enzyme is a more efficient catalyst. In addition, we have determined the structures of three active site mutants which have significantly reduced activity for either the oxidation (His447Asn and His447Gln) or the isomerization (Glu361Gln). Our structural and kinetic data indicate that His447 and Glu361 act as general base catalysts in association with conserved water H2O541 and Asn485. The His447, Glu361, H2O541, and Asn485 hydrogen bond network is conserved among other oxidoreductases. This catalytic tetrad appears to be a structural motif that occurs in flavoenzymes that catalyze the oxidation of unactivated alcohols.  相似文献   

8.
Cysteine 319 in the large subunit of Klebsiella aerogenes urease was identified as an essential catalytic residue based on chemical modification studies (Todd, M.J., and Hausinger, R.P. (1991) J. Biol. Chem. 266, 24327-24331). Through site-directed mutagenesis, this cysteine has been changed independently to alanine, serine, aspartate, and tyrosine. None of these mutations (C319A, C319S, C319D, and C319Y, respectively) affected the size or level of synthesis of the urease subunits as monitored by polyacrylamide gel electrophoresis. The wild type enzyme and each of the mutant proteins was purified and their properties were compared. The C319Y protein possessed no detectable activity, while activity was reduced in C319A, C319S, and C319D to 48, 4.5, and 0.03% of wild type levels under normal assay conditions. All of the active mutants had a small increase in Km when compared to the wild type value. The active mutants displayed a greatly reduced sensitivity to inactivation by iodoacetamide in comparison to the wild type enzyme, confirming our previous assignment of the essential cysteine to this residue based on active site peptide mapping. In contrast to the wild type enzyme, inactivation of the mutant proteins was not affected by the presence of the competitive inhibitor phosphate, suggesting that the remaining slow rate of iodoacetamide inactivation is due to modification away from the active site. The pH dependence of urease activity was substantially altered in the active mutants with C319S and C319D showing a pH optimum near 5.2, and C319A near 6.7, compared to the pH 7.75 optimum of wild type urease. These data are consistent with Cys-319 facilitating catalysis at neutral and basic pH values by participating as a general acid.  相似文献   

9.
Phosphofructokinase from Lactobacillus delbrueckii subspecies bulgaricus (LbPFK) has been reported to be a nonallosteric analogue of phosphofructokinase from Escherichia coli at pH 8.2 [Le Bras et al. (1991) Eur. J. Biochem. 198, 683-687]. A reexamination of the kinetics of this enzyme shows LbPFK to have limited binding affinity toward the allosteric ligands, MgADP and PEP, with dissociation constants of approximately 20 mM for both. Their allosteric effects are observed only at high concentrations of these ligands, with both exhibiting inhibitory effects on substrate binding. No pH dependence was observed for the binding and the influence of MgADP and PEP on the enzyme. To attempt to explain these results, the crystal structure of LbPFK was solved using molecular replacement to 1.86 A resolution. A comparative study of the LbPFK structure with that of phosphofructokinases from E. coli (EcPFK) and Bacillus stearothermophilus (BsPFK) reveals a structure with conserved fold and substrate binding site. The effector binding site, however, shows many differences that could explain the observed decreases in binding affinity for MgADP and PEP in LbPFK as compared to the other two enzymes.  相似文献   

10.
NCOAT is a bifunctional nucleo-cytoplasmic protein with both O-GlcNAcase and histone acetyltransferase domains. The O-GlcNAcase domain catalyzes the removal of O-linked GlcNAc modifications from proteins and we have found that it resides in the N-terminal third of NCOAT. The recognition of the substrate GlcNAc suggests that the O-GlcNAcase is related in structure and catalytic mechanism to chitinases, hexosaminidases and hyaluronidases. These families of glycosidases all possess a catalytic doublet of carboxylate-containing residues, with one providing an acid-base function, and the second acting to orient and use the N-acetyl group of GlcNAc during catalysis. Indeed, we show that the O-GlcNAcase also possesses the catalytic doublet motif shared among these enzymes and that these two essential residues are aspartic acids at positions 175 and 177, respectively, in mouse NCOAT. In addition, a conserved cysteine at 166 and a conserved aspartic acid at 174 were also found to be necessary for fully efficient enzymatic activity. Given this information, we propose that the O-GlcNAcase active site resembles those of the above glycosidases which carry out the hydrolysis of GlcNAc linkages in a substrate-assisted acid-base manner.  相似文献   

11.
Alpha-methylacyl-CoA racemase (Amacr) catalyzes the racemization of alpha-methyl-branched CoA esters. Sequence comparisons have shown that this enzyme is a member of the family III CoA transferases. The mammalian Amacr is involved in bile acid synthesis and branched-chain fatty acid degradation. In human, mutated variants of Amacr have been shown to be associated with disease states. Amino acid sequence alignment of Amacrs and its homologues from various species revealed 26 conserved protic residues, assumed to be potential candidates as catalytic residues. Amacr from Mycobacterium tuberculosis (MCR) was taken as a representative of the racemases. To determine their importance for efficient catalysis, each of these 26 protic residues of MCR was mutated into an alanine, respectively, and the mutated variants were overexpressed in Escherichia coli. It was found that four variants (R91A, H126A, D156A, and E241A) were properly folded but had much decreased catalytic efficiency. Apparently, Arg91, His126, Asp156, and Glu241 are important catalytic residues of MCR. The importance of these residues for catalysis can be rationalized by the 1.8 A resolution crystal structure of MCR, which shows that the catalytic site is at the interface between the large and small domain of two different subunits of the dimeric enzyme. This crystal structure is the first structure of a complete enzyme of the bile acid synthesis pathway. It shows that MCR has unique structural features, not seen in the structures of the sequence related formyl-CoA transferases, suggesting that the family III CoA transferases can be subdivided in at least two classes, being racemases and CoA transferases.  相似文献   

12.
Pyrroloquinoline quinone [4,5-dihydro-4,5-dioxo-1H-pyrrolo[2,3-f]quinoline-2,7,9-tricarboxylic acid (PQQ)] is a bacterial vitamin that serves as a cofactor in numerous alcohol dehydrogenases. Its biosynthesis in Klebsiella pneumoniae is facilitated by six genes, pqqABCDEF, and proceeds by an unknown pathway. The protein encoded by pqqC catalyzes the final step of PQQ formation, which involves a ring closure and an overall eight-electron oxidation of 3a-(2-amino-2-carboxyethyl)-4,5-dioxo-4,5,6,7,8,9-hexahydroquinoline-7,9-dicarboxylic acid (AHQQ) in the absence of a redox-active metal or cofactor. A recent crystal structure has implicated numerous PQQ-PqqC interactions [Magnusson et al. (2004) Proc. Natl. Acad. Sci. U.S.A. 101, 7913-7918]. To investigate the mechanism of the PqqC reaction, the active site residue His84 has been mutated to H84A and H84N, and the kinetic and spectroscopic properties have been compared to each other and the wild-type enzyme using aerobic and anaerobic conditions. Both mutants form PQQ under aerobic conditions with rate constants of 0.09 min-1 and 0.056 min-1 relative to 0.34 min-1 for the wild-type enzyme. In addition to the initial E-AHQQ complex (532-536 nm) and the product E-PQQ complex (346-366 nm), a number of spectral intermediates are observed between 316 and 344 nm. The anaerobic reaction is particularly informative, showing that while mixing of H84N with AHQQ leads to a 344 nm intermediate, this is unable to proceed to a final 318 nm species; by contrast H84A forms the 344 nm species as a precursor to the 318 nm species. In the context of the proposed chemical mechanism for PqqC [Magnusson et al. (2004) Proc. Natl. Acad. Sci. U.S.A. 101, 7913-7918], we assign the 344 nm intermediate to a quinoid species and the 318 nm intermediate to an initial quinol species. The proposed role of H84 is as a proton donor to the oxyanion of the quinoid species such that subsequent C-H bond cleavage can occur to form a monoanionic quinol. In the absence of a proton donor (as occurs in H84N), the normal reaction path is precluded as this would require formation of an unstable, dianionic species. Unlike H84N, H84A appears to be small enough to allow entry of active site water, which is postulated to adopt the role of active site proton donor.  相似文献   

13.
Three arginine residues (Arg-11, Arg-39, Arg-61) are found at the active site of 4-oxalocrotonate tautomerase in the X-ray structure of the affinity-labeled enzyme [Taylor, A. B., Czerwinski, R. M., Johnson, R. M., Jr., Whitman, C. P., and Hackert, M. L. (1998) Biochemistry 37, 14692-14700]. The catalytic roles of these arginines were examined by mutagenesis, kinetic, and heteronuclear NMR studies. With a 1,6-dicarboxylate substrate (2-hydroxymuconate), the R61A mutation showed no kinetic effects, while the R11A mutation decreased k(cat) 88-fold and increased K(m) 8.6-fold, suggesting both binding and catalytic roles for Arg-11. With a 1-monocarboxylate substrate (2-hydroxy-2,4-pentadienoate), no kinetic effects of the R11A mutation were found, indicating that Arg-11 interacts with the 6-carboxylate of the substrate. The stereoselectivity of the R11A-catalyzed protonation at C-5 of the dicarboxylate substrate decreased, while the stereoselectivity of protonation at C-3 of the monocarboxylate substrate increased in comparison with wild-type 4-OT, indicating the importance of Arg-11 in properly orienting the dicarboxylate substrate by interacting with the charged 6-carboxylate group. With 2-hydroxymuconate, the R39A and R39Q mutations decreased k(cat) by 125- and 389-fold and increased K(m) by 1.5- and 2.6-fold, respectively, suggesting a largely catalytic role for Arg-39. The activity of the R11A/R39A double mutant was at least 10(4)-fold lower than that of the wild-type enzyme, indicating approximate additivity of the effects of the two arginine mutants on k(cat). For both R11A and R39Q, 2D (1)H-(15)N HSQC and 3D (1)H-(15)N NOESY-HSQC spectra showed chemical shift changes mainly near the mutated residues, indicating otherwise intact protein structures. The changes in the R39Q mutant were mainly in the beta-hairpin from residues 50 to 57 which covers the active site. HSQC titration of R11A with the substrate analogue cis, cis-muconate yielded a K(d) of 22 mM, 37-fold greater than the K(d) found with wild-type 4-OT (0.6 mM). With the R39Q mutant, cis, cis-muconate showed negative cooperativity in active site binding with two K(d) values, 3.5 and 29 mM. This observation together with the low K(m) of 2-hydroxymuconate (0.47 mM) suggests that only the tight binding sites function catalytically in the R39Q mutant. The (15)Nepsilon resonances of all six Arg residues of 4-OT were assigned, and the assignments of Arg-11, -39, and -61 were confirmed by mutagenesis. The binding of cis,cis-muconate to wild-type 4-OT upshifts Arg-11 Nepsilon (by 0.05 ppm) and downshifts Arg-39 Nepsilon (by 1.19 ppm), indicating differing electronic delocalizations in the guanidinium groups. A mechanism is proposed in which Arg-11 interacts with the 6-carboxylate of the substrate to facilitate both substrate binding and catalysis and Arg-39 interacts with the 1-carboxylate and the 2-keto group of the substrate to promote carbonyl polarization and catalysis, while Pro-1 transfers protons from C-3 to C-5. This mechanism, together with the effects of mutations of catalytic residues on k(cat), provides a quantitative explanation of the 10(7)-fold catalytic power of 4-OT. Despite its presence in the active site in the crystal structure of the affinity-labeled enzyme, Arg-61 does not play a significant role in either substrate binding or catalysis.  相似文献   

14.
The urease proteins of the jack bean (Canavalia ensiformis) and Helicobacter pylori are similar in molecular mass when separated by non-denaturing gradient polyacrylamide gel electrophoresis, both having three main forms. The molecular mass of their major protein form is within the range 440-480 kDa with the other two lesser forms at 230-260 kDa and 660-740 kDa. These forms are all urease active; however, significant kinetic differences exist between the H. pylori and jack bean ureases. Jack bean urease has a single pH optimum at 7.4, whereas H. pylori urease has two pH optima of 4.6 and 8.2 in barbitone and phosphate buffers that were capable of spanning the pH range 3 to 10. The H. pylori Km was 0.6 mM at pH 4.6 and 1.0 mM at pH 8.2 in barbitone buffer, greater than 10.0 mM, and 1.1 mM respectively in phosphate buffer and also greater than 10.0 mM in Tris.HCl at pH 8.2. By comparison, the jack bean urease had a Km of 1.3 mM in Tris.HCl under our experimental conditions. The findings show that the urease activity of H. pylori was inhibited at the pH optimum of 4.6 in the phosphate buffer, but not in the barbitone buffer. This was shown to be due to competitive inhibition by the sodium and potassium ions in the phosphate buffer, not the phosphate ions as suggested earlier. Jack bean urease activity was similarly inhibited by phosphate buffer but again due to the effect of sodium and potassium ions.  相似文献   

15.
Xanthine/α-ketoglutarate (αKG) dioxygenase (XanA) is a non-heme mononuclear FeII enzyme that decarboxylates αKG to succinate and CO2 while hydroxylating xanthine to generate uric acid. In the absence of a XanA crystal structure, a homology model was used to target several putative active site residues for mutagenesis. Wild-type XanA and ten enzyme variants were purified from recombinant Escherichia coli cells and characterized. The H149A and D151A variants were inactive and the H340A variant exhibited only 0.17% of the wild-type enzyme activity, consistent with the proposed role of His149, Asp151, and His340 as Fe ligands. The K122A variant led to a 2-fold increase in the Kd of αKG as measured by fluorescence quenching analysis, in agreement with Lys122 acting to stabilize the binding of αKG. The N358A variant exhibited a 23-fold decrease in kcat/Km compared to wild-type XanA, pointing to a key role of Asn358 in catalysis. 9-Methylxanthine was exploited as an alternate substrate, and the C357A, E137A, and D138A variants were found to exhibit relatively enhanced activity consistent with Cys357, Glu137, and Asp138 being proximal to N-9 or involved in its proper positioning. 6,8-Dihydroxypurine was identified as a slow-binding competitive inhibitor of XanA, and significant decreases (E137A and D138A) or increases (Q356A and N358A) in of the variants were interpreted in terms of distinct interactions between this compound and the corresponding active site side chains. Further support for Cys357 residing at the active site was obtained using thiol-specific reagents that inactivated wild-type enzyme (with partial protection by substrate), whereas the C357A variant was resistant to these reagents. The Q101A, Q356A, and C357A variants showed elevated ferroxidase activity in the absence of substrates, pointing to the presence of the corresponding side chains at the active site. These results confirm most aspects of the homology model and provide additional insight into the enzyme reactivity.  相似文献   

16.
Chondroitinase B from Flavobacterium heparinum is the only known lyase that cleaves the glycosaminoglycan, dermatan sulfate (DS), as its sole substrate. A recent co-crystal structure of chondroitinase B with a disaccharide product of DS depolymerization has provided some insight into the location of the active site and suggested potential roles of some active site residues in substrate binding and catalysis. However, this co-crystal structure was not representative of the actual enzyme-substrate complex, because the disaccharide product did not have the right length or the chemical structure of the minimal substrate (tetrasaccharide) involved in catalysis. Therefore, only a limited picture of the functional role of active site residues in DS depolymerization was presented in previous structural studies. In this study, by docking a DS tetrasaccharide into the proposed active site of the enzyme, we have identified novel roles of specific active site amino acids in the catalytic function of chondroitinase B. Our conformational analysis also revealed a unique, symmetrical arrangement of active site amino acids that may impinge on the catalytic mechanism of action of chondroitinase B. The catalytic residues Lys-250, Arg-271, His-272, and Glu-333 along with the substrate binding residues Arg-363 and Arg-364 were mutated using site-directed mutagenesis, and the kinetics and product profile of each mutant were compared with recombinant chondroitinase B. Mutating Lys-250 to alanine resulted in inactivation of the enzyme, potentially attributable to the role of the residue in stabilizing the carbanion intermediate formed during enzymatic catalysis. The His-272 and Glu-333 mutants showed diminished enzymatic activity that could be indicative of a possible role for one or both residues in the abstraction of the C-5 proton from the galactosamine. In addition, the Arg-364 mutant had an altered product profile after exhaustive digestion of DS, suggesting a role for this residue in defining the substrate specificity of chondroitinase B.  相似文献   

17.
We examined several compounds for their mechanisms of inhibition with the nickel-containing active site of homogeneous Klebsiella aerogenes urease. Thiolate anions competitively inhibit urease and directly interact with the metallocenter, as shown by the pH dependence of inhibition and by UV-visible absorbance spectroscopic studies. Cysteamine, which possesses a cationic beta-amino group, exhibited a high affinity for urease (Ki = 5 microM), whereas thiolates containing anionic carboxyl groups were uniformly poor inhibitors. Phosphate monoanion competitively inhibits a protonated form of urease with a pKa of less than 5. Both the thiolate and phosphate inhibition results are consistent with charge repulsion by an anionic group in the urease active site. Acetohydroxamic acid (AHA) was shown to be a slow-binding competitive inhibitor of urease. This compound forms an initial E.AHA complex which then undergoes a slow transformation to yield an E.AHA* complex; the overall dissociation constant of AHA is 2.6 microM. Phenylphosphorodiamidate, also shown to be a slow-binding competitive inhibitor, possesses an overall dissociation constant of 94 pM. The tight binding of phenylphosphorodiamidate was exploited to demonstrate the presence of two active sites per enzyme molecule. Urease contains 4 mol of nickel/mol enzyme, hence there are two nickel ions/catalytic unit. Each of the two slow-binding inhibitors are proposed to form complexes in which the inhibitor bridges the two active site nickel ions. The inhibition results obtained for K. aerogenes urease are compared with inhibition studies of other ureases and are interpreted in terms of a model for catalysis proposed for the jack bean enzyme (Dixon, N.E., Riddles, P.W., Gazzola, C., Blakely, R.L., and Zerner, B. (1980) Can. J. Biochem. 58, 1335-1344).  相似文献   

18.
19.
DNA sequence analysis of the stuctural urease genes from Staphylococcus xylosus revealed that three enzyme subunits are encoded in the order of 11000, 15400 and 61000 (mol. mass), which correspond to the single polypeptide chain of jack bean urease (90800). Comparing the deduced amino acid sequence of S. xylosus urease with the amino acid sequence of jack bean urease an overall portion of 56% identical residues was found. For S. xylosus urease a subunit structure of ()4 was proposed, based on the comparison of the deduced amino acid content of the enzyme subunits with the total amino acid content of the purified enzyme. The staphylococcal enzyme contained no cysteine, as deduced from DNA sequence and confirmed by the determination of the total amino acid content in the purified enzyme. Instead of cysteine, known to be catalytically essential in the plant enzyme, and conserved among all bacterial ureases analyzed so far, threonine was found in S. xylosus. This amino acid-exchange was located within a highly conserved domain of 17 amino acids, supposed to be part of the active site. Sequence analysis of the respective region of Staphylococcus saprophyticus urease showed that it also contains threonine instead of cysteine. In contrast to jack bean urease S. xylosus urease was not affected by the SH-group inhibitor dipyridyl disulfide but was completely inhibited by the serine protease inhibitor phenylmethanesulfonyl fluoride. The presented results indicated that in these staphylococcal strains urea hydrolysis might function in a manner similar to the peptide bond cleavage by chymotrypsin.Abbreviations AA amino acid - ATZ anilino thiazolinone - DPDS dipyridyl disulfide - Kb kilobase pairs - PITC phenylisothiocyanate - PTH phenylthiohydantoin - PMSF phenylmethanesulfonyl fluoride  相似文献   

20.
The E. coli isozyme of gamma-aminobutyrate aminotransferase (GABA-AT) is a tetrameric pyridoxal phosphate-dependent enzyme that catalyzes transamination between primary amines and alpha-keto acids. The roles of the active site residues V241, E211, and I50 in the GABA-AT mechanism have been probed by site-directed mutagenesis. The beta-branched side chain of V241 facilitates formation of external aldimine intermediates with primary amine substrates, while E211 provides charge compensation of R398 selectively in the primary amine half-reaction and I50 forms a hydrophobic lid at the top of the substrate binding site. The structures of the I50Q, V241A, and E211S mutants were solved by X-ray crystallography to resolutions of 2.1, 2.5, and 2.52 A, respectively. The structure of GABA-AT is similar in overall fold and active site structure to that of dialkylglycine decarboxylase, which catalyzes both transamination and decarboxylation half-reactions in its normal catalytic cycle. Therefore, an attempt was made to convert GABA-AT into a decarboxylation-dependent aminotransferase similar to dialkylglycine decarboxylase by systematic mutation of E. coli GABA-AT active site residues. Two of the twelve mutants presented, E211S/I50G/C77K and E211S/I50H/V80D, have approximately 10-fold higher decarboxylation activities than the wild-type enzyme, and the E211S/I50H/V80D has formally changed the reaction specificity to that of a decarboxylase.  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