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.
Li JJ  Li C  Blindauer CA  Bugg TD 《Biochemistry》2006,45(41):12461-12469
C-C hydrolase enzymes MhpC and BphD catalyze the hydrolytic C-C cleavage of meta-ring fission intermediates on the Escherichia coli phenylpropionic acid and Burkholderia xenovorans LB400 biphenyl degradation pathways and are both members of the alpha/beta-hydrolase family containing a Ser-His-Asp catalytic triad. The catalytic mechanism of this family of enzymes is thought to proceed via a gem-diol reaction intermediate, which has not been observed directly. Site-directed single mutants of BphD in which catalytic residues His-265 and Ser-112 were replaced with Ala were found to possess 10(4)-fold reduced k(cat) values, and in each case, the C-C cleavage step was shown by pre-steady-state kinetic analysis to be rate-limiting. The processing of a 6-(13)C-labeled aryl-containing substrate by these H265A or S112A mutant BphD enzymes was monitored directly by (13)C NMR spectroscopy. A new line-broadened signal was observed at 128 ppm for each enzyme, corresponding to the proposed gem-diol reaction intermediate, over a time scale of 1-24 h. A similar signal was observed upon incubation of the (13)C-labeled substrate with an H114A MhpC mutant, which is able to accept the 6-phenyl-containing substrate, on a shorter time scale. The direct observation of a gem-diol intermediate provides further evidence that supports a general base mechanism for this family of enzymes.  相似文献   

2.
Choline kinase catalyzes the phosphorylation of choline by ATP, the first committed step in the CDP-choline pathway for phosphatidylcholine biosynthesis. To begin to elucidate the mechanism of catalysis by this enzyme, choline kinase A-2 from Caenorhabditis elegans was analyzed by systematic mutagenesis of highly conserved residues followed by analysis of kinetic and structural parameters. Specifically, mutants were analyzed with respect to K(m) and k(cat) values for each substrate and Mg(2+), inhibitory constants for Mg(2+) and Ca(2+), secondary structure as monitored by circular dichroism, and sensitivity to unfolding in guanidinium hydrochloride. The most severe impairment of catalysis occurred with the modification of Asp-255 and Asn-260, which are located in the conserved Brenner's phosphotransferase motif, and Asp-301 and Glu-303, in the signature choline kinase motif. For example, mutation of Asp-255 or Asp-301 to Ala eliminated detectable catalytic activity, and mutation of Asn-260 and Glu-303 to Ala decreased k(cat) by 300- and 10-fold, respectively. Additionally, the K(m) for Mg(2+) for mutants N260A and E303A was approximately 30-fold higher than that of wild type. Several other residues (Ser-86, Arg-111, Glu-125, and Trp-387) were identified as being important: Catalytic efficiencies (k(cat)/K(m)) for the enzymes in which these residues were mutated to Ala were reduced to 2-25% of wild type. The high degree of structural similarity among choline kinase A-2, aminoglycoside phosphotransferases, and protein kinases, together with the results from this mutational analysis, indicates it is likely that these conserved residues are located at the catalytic core of choline kinase.  相似文献   

3.
Ruzzini AC  Horsman GP  Eltis LD 《Biochemistry》2012,51(29):5831-5840
meta-Cleavage product (MCP) hydrolases catalyze C-C bond fission in the aerobic catabolism of aromatic compounds by bacteria. These enzymes utilize a Ser-His-Asp triad to catalyze hydrolysis via an acyl-enzyme intermediate. BphD, which catalyzes the hydrolysis of 2-hydroxy-6-oxo-6-phenylhexa-2,4-dienoic acid (HOPDA) in biphenyl degradation, catalyzed the hydrolysis of an ester analogue, p-nitrophenyl benzoate (pNPB), with a k(cat) value (6.3 ± 0.5 s(-1)) similar to that of HOPDA (6.5 ± 0.5 s(-1)). Consistent with the breakdown of a shared intermediate, product analyses revealed that BphD catalyzed the methanolysis of both HOPDA and pNPB, partitioning the products to benzoic acid and methyl benzoate in similar ratios. Turnover of HOPDA was accelerated up to 4-fold in the presence of short, primary alcohols (methanol > ethanol > n-propanol), suggesting that deacylation is rate-limiting during catalysis. In the steady-state hydrolysis of HOPDA, k(cat)/K(m) values were independent of methanol concentration, while both k(cat) and K(m) values increased with methanol concentration. This result was consistent with a simple model of nucleophilic catalysis. Although the enzyme could not be saturated with pNPB at methanol concentrations of >250 mM, k(obs) values from the steady-state turnover of pNPB at low methanol concentrations were also consistent with a nucleophilic mechanism of catalysis. Finally, transient-state kinetic analysis of pNPB hydrolysis by BphD variants established that substitution of the catalytic His reduced the rate of acylation by more than 3 orders of magnitude. This suggests that for pNPB hydrolysis, the serine nucleophile is activated by the His-Asp dyad. In contrast, rapid acylation of the H265Q variant during C-C bond cleavage suggests that the serinate forms via a substrate-assisted mechanism. Overall, the data indicate that ester hydrolysis proceeds via the same acyl-enzyme intermediate as that of the physiological substrate but that the serine nucleophile is activated via a different mechanism.  相似文献   

4.
The Met-542 residue of β-galactosidase is important for the enzyme's activity because it acts as a guide for the movement of the benzyl side chain of Phe-601 between two stable positions. This movement occurs in concert with an important conformational change (open vs. closed) of an active site loop (residues 794-803). Phe-601 and Arg-599, which interact with each other via the π electrons of Phe-601 and the guanidium cation of Arg-599, move out of their normal positions and become disordered when Met-542 is replaced by an Ala residue because of the loss of the guide. Since the backbone carbonyl of Phe-601 is a ligand for Na(+), the Na(+) also moves out of its normal position and becomes disordered; the Na(+) binds about 120 times more poorly. In turn, two other Na(+) ligands, Asn-604 and Asp-201, become disordered. A substrate analog (IPTG) restored Arg-599, Phe-601, and Na(+) to their normal open-loop positions, whereas a transition state analog d-galactonolactone) restored them to their normal closed-loop positions. These compounds also restored order to Phe-601, Asn-604, Asp-201, and Na(+). Binding energy was, however, necessary to restore structure and order. The K(s) values of oNPG and pNPG and the competitive K(i) values of substrate analogs were 90-250 times higher than with native enzyme, whereas the competitive K(i) values of transition state analogs were ~3.5-10 times higher. Because of this, the E?S energy level is raised more than the E?transition state energy level and less activation energy is needed for galactosylation. The galactosylation rates (k?) of M542A-β-galactosidase therefore increase. However, the rate of degalactosylation (k?) decreased because the E?transition state complex is less stable.  相似文献   

5.
D H Kim  G H Nam  D S Jang  G Choi  S Joo  J S Kim  B H Oh  K Y Choi 《Biochemistry》1999,38(42):13810-13819
The aromatic residues Phe-54, Phe-82, and Trp-116 in the hydrophobic substrate-binding pocket of Delta(5)-3-ketosteroid isomerase from Pseudomonas putida biotype B have been characterized in their roles in steroid binding and catalysis. Kinetic and equilibrium binding analyses were carried out for the mutant enzymes with the substitutions Phe-54 --> Ala or Leu, Phe-82 --> Ala or Leu, and Trp-116 --> Ala, Phe, or Tyr. The removal of their bulky, aromatic side chains at any of these three positions results in reduced k(cat), particularly when Phe-82 or Trp-116 is replaced by Ala. The results are consistent with the binding interactions of the aromatic residues with the bound steroid contributing to catalysis. All the mutations except the F82A mutation increase K(m); the F82A mutation decreases K(m) by ca. 3-fold, suggesting a possibility that the phenyl ring at position 82 might be unfavorable for substrate binding. The K(D) values for d-equilenin, an intermediate analogue, suggest that a space-filling hydrophobic side chain at position 54, a phenyl ring at position 82, and a nonpolar aromatic or small side chain at position 116 might be favorable for binding the reaction intermediate. In contrast to the increased K(D) for equilenin, the enzymes with any substitutions at positions 54 and 116 display a decreased K(D) for 19-nortestosterone, a product analogue, indicating that Phe-54 and Trp-116 might be unfavorable for product binding. The crystal structure of F82A determined to 2.1-A resolution reveals that Phe-82 is important for maintaining the active site geometry. Taken together, our results demonstrate that Phe-54, Phe-82, and Trp-116 contribute differentially to the stabilization of steroid species including substrate, intermediate, and product.  相似文献   

6.
trans-3-Chloroacrylic acid dehalogenase (CaaD) catalyzes the hydrolytic dehalogenation of trans-3-haloacrylates to yield malonate semialdehyde by a mechanism utilizing βPro-1, αArg-8, αArg-11, and αGlu-52. These residues are implicated in a promiscuous hydratase activity where 2-oxo-3-pentynoate is processed to acetopyruvate. The roles of three nearby residues (βAsn-39, αPhe-39, and αPhe-50) are unexplored. Mutants were constructed at these positions (βN39A, αF39A, αF39T, αF50A and αF50Y) and kinetic parameters determined along with those of the αR8K and αR11K mutants. Analysis indicates that αArg-8, αArg-11, and βAsn-39 are critical for dehalogenase activity whereas αArg-11 and αPhe-50 are critical for hydratase activity. Docking studies suggest structural bases for these observations.  相似文献   

7.
Structural and kinetic data show that Arg-599 of β-galactosidase plays an important role in anchoring the "open" conformations of both Phe-601 and an active-site loop (residues 794-803). When alanine was substituted for Arg-599, the conformations of Phe-601 and the loop shifted towards the "closed" positions because interactions with the guanidinium side chain were lost. Also, Phe-601, the loop, and Na+, which is ligated by the backbone carbonyl of Phe-601, lost structural order, as indicated by large B-factors. IPTG, a substrate analog, restored the conformations of Phe-601 and the loop of R599A-β-galactosidase to the open state found with IPTG-complexed native enzyme and partially reinstated order. ?-Galactonolactone, a transition state analog, restored the closed conformations of R599A-β-galactosidase to those found with ?-galactonolactone-complexed native enzyme and completely re-established the order. Substrates and substrate analogs bound R599A-β-galactosidase with less affinity because the closed conformation does not allow substrate binding and extra energy is required for Phe-601 and the loop to open. In contrast, transition state analog binding, which occurs best when the loop is closed, was several-fold better. The higher energy level of the enzyme?substrate complex and the lower energy level of the first transition state means that less activation energy is needed to form the first transition state and thus the rate of the first catalytic step (k2) increased substantially. The rate of the second catalytic step (k3) decreased, likely because the covalent form is more stabilized than the second transition state when Phe-601 and the loop are closed. The importance of the guanidinium group of Arg-599 was confirmed by restoration of conformation, order, and activity by guanidinium ions.  相似文献   

8.
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   

9.
DypB from Rhodococcus jostii RHA1 is a bacterial dye-decolorizing peroxidase (DyP) that oxidizes lignin and Mn(II). Three residues interact with the iron-bound solvent species in ferric DypB: Asn-246 and the conserved Asp-153 and Arg-244. Substitution of either Asp-153 or Asn-246 with alanine minimally affected the second order rate constant for Compound I formation (k(1) ~ 10(5) M(-1)s(-1)) and the specificity constant (k(cat)/K(m)) for H(2)O(2). Even in the D153A/N246A double variant, these values were reduced less than 30-fold. However, these substitutions dramatically reduced the stability of Compound I (t(1/2) ~ 0.13 s) as compared with the wild-type enzyme (540 s). By contrast, substitution of Arg-244 with leucine abolished the peroxidase activity, and heme iron of the variant showed a pH-dependent transition from high spin (pH 5) to low spin (pH 8.5). Two variants were designed to mimic the plant peroxidase active site: D153H, which was more than an order of magnitude less reactive with H(2)O(2), and N246H, which had no detectable peroxidase activity. X-ray crystallographic studies revealed that structural changes in the variants are confined to the distal heme environment. The data establish an essential role for Arg-244 in Compound I formation in DypB, possibly through charge stabilization and proton transfer. The principle roles of Asp-153 and Asn-246 appear to be in modulating the subsequent reactivity of Compound I. These results expand the range of residues known to catalyze Compound I formation in heme peroxidases.  相似文献   

10.
Thromboxane A2 synthase (TXAS) is a member of the cytochrome P450 superfamily and catalyzes an isomerization reaction that converts prostaglandin H2 to thromboxane A2. As a step toward understanding the structure/function relationships of TXAS, we mutated amino acid residues predicted to bind the propionate groups of A- and D-pyrrole rings of the heme. These mutations at each of these residues (Asn-110, Trp-133, Arg-137, Arg-413, and Arg-478) resulted in altered heme binding, as evidenced by perturbation of the absorption spectra and EPR. The mutations, although causing no significant changes in the secondary structure of the proteins, induced tertiary structural changes that led to increased susceptibility to trypsin digestion and alteration of the intrinsic protein fluorescence. Moreover, these mutant proteins lost their binding affinity to the substrate analog, had a lower heme content and retained less than 5% of the wild-type catalytic activity. However, mutations at the neighboring amino acid of the aforementioned residues yielded mutant proteins retaining the biochemical and biophysical properties of the wild type TXAS. Aligning the TXAS sequence with the structurally known P450s, we proposed that in TXAS the A-ring propionate of the heme is hydrogen bonded to Asn-110, Arg-413, and Arg-478, whereas D-ring propionate is hydrogen bonded to Trp-133 and Arg-137. Furthermore, both A- and D-ring propionates bulge away from the heme plane and both lie on the proximal face of heme plane, a structure similar to P450terp.  相似文献   

11.
Aminopeptidase A (EC 3.4.11.7, APA) is a 160 kDa membrane-bound zinc enzyme that contains the HEXXH consensus sequence found in members of the zinc metalloprotease family, the zincins. In addition, the monozinc aminopeptidases are characterized by another conserved motif, GXMEN, the glutamate residue of which has been shown to be implicated in the exopeptidase specificity of aminopeptidase A [Vazeux G. (1998) Biochem. J. 334, 407-413]. In carboxypeptidase A (EC 3.4.17.1, CPA), the exopeptidase specificity is conferred by an arginine residue (Arg-145) and an asparagine residue (Asn-144). Thus, we hypothesized that Asn-353 of the GXMEN motif in APA plays a similar role to Asn-144 in CPA and contributes to the exopeptidase specificity of APA. We investigated the functional role of Asn-353 in APA by substituting this residue with a glutamine (Gln-353), an alanine (Ala-353) or an aspartate (Asp-353) residue by site-directed mutagenesis. Expression of wild-type and mutated APAs revealed that Gln-353 and Ala-353 are similarly routed and glycosylated to the wild-type APA, whereas Asp-353 is trapped intracellularly and partially glycosylated. Kinetic studies, using alpha-L-glutamyl-beta-naphthylamide (GluNA) as a substrate showed that the K(m) values of the mutants Gln-353 and Ala-353 were increased 11- and 8-fold, respectively, whereas the k(cat) values were decreased (2-fold) resulting in a 24- and 14-fold reduction in cleavage efficiency. When alpha-L-aspartyl-beta-naphthylamide or angiotensin II were used as substrates, the mutations had a greater effect on k(cat), leading to a similar decrease in cleavage efficiencies as that observed with GluNA. We then measured the inhibitory potencies of several classes of inhibitors, glutamate thiol, glutamine thiol and two isomers (L- or D-) of glutamate phosphonate to explore the functional role of Asn-353. The data indicate that Asn-353 is critical for the integrity and catalytic activity of APA. This residue is involved in substrate binding via interactions with the free N-terminal part and with the P1 carboxylate side chain of the substrate. In conclusion, Asn-353 of the GXMEN motif, together with Glu-352, contributes to the exopeptidase specificity of APA and plays an equivalent role to Asn-144 in CPA.  相似文献   

12.
The tetradecapeptide Ac-D-F-L-A-E-G-G-G-V-R-G-P-R-V-OMe, which mimics residues 7f-20f of the A alpha-chain of human fibrinogen, has been co-crystallized with bovine thrombin from ammonium sulfate solutions in space group P2(1) with unit cell dimensions of a = 83.0 A, b = 89.4 A, c = 99.3 A, and beta = 106.6 degrees. Three crystallographically independent complexes were located in the asymmetric unit by molecular replacement using the native bovine thrombin structure as a model. The standard crystallographic R-factor is 0.167 at 2.3-A resolution. Excellent electron density could be traced for the decapeptide, beginning with Asp-7f and ending with Arg-16f in the active site of thrombin; the remaining 4 residues, which have been cleaved from the tetradecapeptide at the Arg-16f/Gly-17f bond, are not seen. Residues 7f-11f at the NH2 terminus of the peptide form a single turn of alpha-helix that is connected by Gly-12f, which has a positive phi angle, to an extended chain containing residues 13f-16f. The major specific interactions between the peptide and thrombin are 1) a hydrophobic cage formed by residues Tyr-60A, Trp-60D, Leu-99, Ile-174, Trp-215, Leu-9f, Gly-13f, and Val-15f that surrounds Phe-8f; 2) a hydrogen bond linking Phe-8f NH to Lys-97 O;3) a salt link between Glu-11f and Arg-173; 4) two antiparallel beta-sheet hydrogen bonds between Gly-14f and Gly-216; and 5) the insertion of Arg-16f into the specificity pocket. Binding of the peptide is accompanied by a considerable shift in two of the loops near the active site relative to human D-phenyl-L-prolyl-L-arginyl chloromethyl ketone (PPACK)-thrombin.  相似文献   

13.
The Streptomyces sp. beta-glucosidase (Bgl3) is a retaining glycosidase that belongs to family 1 glycosyl hydrolases. Steady-state kinetics with p-nitrophenyl beta-D-glycosides revealed that the highest k(cat)/K(M) values are obtained with glucoside (with strong substrate inhibition) and fucoside (with no substrate inhibition) substrates and that Bgl3 has 10-fold glucosidase over galactosidase activity. Reactivity studies by means of a Hammett analysis using a series of substituted aryl beta-glucosides gave a biphasic plot log k(cat) vs pK(a) of the phenol aglycon: a linear region with a slope of beta(lg) = -0.8 for the less reactive substrates (pK(a) > 8) and no significant dependence for activated substrates (pK(a) < 8). Thus, according to the two-step mechanism of retaining glycosidases, formation of the glycosyl-enzyme intermediate is rate limiting for the former substrates, while hydrolysis of the intermediate is for the latter. To identify key catalytic residues and on the basis of sequence similarity to other family 1 beta-glucosidases, glutamic acids 178 and 383 were changed to glutamine and alanine by site-directed mutagenesis. Mutation of Glu178 to Gln and Ala yielded enzymes with 250- and 3500-fold reduction in their catalytic efficiencies, whereas larger reduction (10(5)-10(6)-fold) were obtained for mutants at Glu383. The functional role of both residues was probed by a chemical rescue methodology based on activation of the inactive Ala mutants by azide as exogenous nucleophile. The E178A mutant yielded the beta-glucosyl azide adduct (by (1)H NMR) with a 200-fold increase on k(cat) for the 2,4-dinitrophenyl glucoside but constant k(cat)/K(M) on azide concentration. On the other hand, the E383A mutant with the same substrate gave the alpha-glucosyl azide product and a 100-fold increase in k(cat) at 1 M azide. In conclusion, Glu178 is the general acid/base catalyst and Glu383 the catalytic nucleophile. The results presented here indicate that Bgl3 beta-glucosidase displays kinetic and mechanistic properties similar to other family 1 enzymes analyzed so far. Subtle differences in behavior would lie in the fine and specific architecture of their respective active sites.  相似文献   

14.
We used site-directed mutagenesis to determine the minimum number of PDGF B residues needed to convert PDGF A to a potently transforming PDGF B-like molecule. Substitution of two PDGF B subdomains, 106-115 and 135-144, were found to be critical. These substitutions were sufficient to broaden the ability of PDGF A to activate beta as well as alpha platelet-derived growth factor (PDGF) receptors and increase its transforming efficiency to that of PDGF B. Within subdomain I, either PDGF B residues Arg-109 and Asn-115 or Arg-109, Leu-110, and Arg-113, in combination with subdomain II PDGF B residues Asn-136, Arg-137, and Arg-142 were identified as being essential. Those mutants with transforming ability comparable with PDGF B showed significantly lower efficiencies of beta receptor triggering. Thus, our studies identify a small number of PDGF B amino acids indispensable for beta PDGF receptor interaction and suggest that a low level of beta PDGF receptor activation is sufficient to dramatically increase PDGF transforming efficiency in NIH 3T3 cells.  相似文献   

15.
The recent elucidation of crystal structures of a bacterial member of the NCS1 family, the Mhp1 benzyl-hydantoin permease from Microbacterium liquefaciens, allowed us to construct and validate a three-dimensional model of the Aspergillus nidulans purine-cytosine/H+ FcyB symporter. The model consists of 12 transmembrane α-helical, segments (TMSs) and cytoplasmic N- and C-tails. A distinct core of 10 TMSs is made of two intertwined inverted repeats (TMS1–5 and TMS6–10) that are followed by two additional TMSs. TMS1, TMS3, TMS6, and TMS8 form an open cavity that is predicted to host the substrate binding site. Based on primary sequence alignment, three-dimensional topology, and substrate docking, we identified five residues as potentially essential for substrate binding in FcyB; Ser-85 (TMS1), Trp-159, Asn-163 (TMS3), Trp-259 (TMS6), and Asn-354 (TMS8). To validate the role of these and other putatively critical residues, we performed a systematic functional analysis of relevant mutants. We show that the proposed substrate binding residues, plus Asn-350, Asn-351, and Pro-353 are irreplaceable for FcyB function. Among these residues, Ser-85, Asn-163, Asn-350, Asn-351, and Asn-354 are critical for determining the substrate binding affinity and/or the specificity of FcyB. Our results suggest that Ser-85, Asn-163, and Asn-354 directly interact with substrates, Trp-159 and Trp-259 stabilize binding through π-π stacking interactions, and Pro-353 affects the local architecture of substrate binding site, whereas Asn-350 and Asn-351 probably affect substrate binding indirectly. Our work is the first systematic approach to address structure-function-specificity relationships in a eukaryotic member of NCS1 family by combining genetic and computational approaches.  相似文献   

16.
The NH(2) terminus (residues 1-59) of streptokinase (SK) is a molecular switch that permits fibrin-independent plasminogen activation. Targeted mutations were made in recombinant (r) SK1-59 to identify structural interactions required for this process. Mutagenesis established the functional roles of Phe-37and Glu-39, which were projected to interact with microplasmin in the activator complex. Mutation of Leu-42 (rSK1-59(L42A)), a conserved residue in the SK fibronectin motif that lacks interactions with microplasmin, strongly reduced plasminogen activation (k(cat) decreased 50-fold) but not amidolysis (k(cat) decreased 1.5-fold). Otherwise rSK1-59(L42A) and native rSK1-59 were indistinguishable in several parameters. Both displayed saturable and specific binding to Glu-plasminogen or the remaining SK fragment (rSKDelta59). Similarly rSK1-59 and rSK1-59(L42A) bound simultaneously to two different plasminogen molecules, indicating that both plasminogen binding sites were intact. However, when bound to SKDelta59, rSK1-59(L42A) was less effective than rSK1-59 in restructuring the native conformation of the SK A domain, as detected by conformation-dependent monoclonal antibodies. In the light of previous studies, these data provide evidence that SK1-59 contributes to fibrin-independent plasminogen activation through 1) intermolecular interactions with the plasmin in the activator complex, 2) binding interactions with the plasminogen substrate, and 3) intramolecular interactions that structure the A domain of SK for Pg substrate processing.  相似文献   

17.
Sphingomonas wittichii RW1 degrades chlorinated dibenzofurans and dibenzo-p-dioxins via meta cleavage. We used inverse PCR to amplify dxnB2, a gene encoding one of three meta-cleavage product (MCP) hydrolases identified in the organism that are homologues of BphD involved in biphenyl catabolism. Purified DxnB2 catalyzed the hydrolysis of 8-OH 2-hydroxy-6-oxo-6-phenylhexa-2,4-dienoate (HOPDA) approximately six times faster than for HOPDA at saturating substrate concentrations. Moreover, the specificity of DxnB2 for HOPDA (k(cat)/K(m) = 1.2 x 10(7) M(-1) s(-1)) was about half that of the BphDs of Burkholderia xenovorans LB400 and Rhodococcus globerulus P6, two potent polychlorinated biphenyl (PCB)-degrading strains. Interestingly, DxnB2 transformed 3-Cl and 4-OH HOPDAs, compounds that inhibit the BphDs and limit PCB degradation. DxnB2 had a higher specificity for 9-Cl HOPDA than for HOPDA but a lower specificity for 8-Cl HOPDA (k(cat)/K(m) = 1.7 x 10(6) M(-1) s(-1)), the chlorinated analog of 8-OH HOPDA produced during dibenzofuran catabolism. Phylogenetic analyses based on structure-guided sequence alignment revealed that DxnB2 belongs to a previously unrecognized class of MCP hydrolases, evolutionarily divergent from the BphDs although the physiological substrates of both enzyme types are HOPDAs. However, both classes of enzymes have mainly small hydrophobic residues lining the subsite that binds the C-6 phenyl of HOPDA, in contrast to the bulky hydrophobic residues (Phe106, Phe135, Trp150, and Phe197) found in the class II enzymes that prefer substrates possessing a C-6 alkyl. Thr196 and/or Asn203 appears to be an important determinant of specificity for DxnB2, potentially forming hydrogen bonds with the 8-OH substituent. This study demonstrates that the substrate specificities of evolutionarily divergent hydrolases may be useful for degrading mixtures of pollutants, such as PCBs.  相似文献   

18.
碳-碳水解酶(C-C水解酶)作为α/β水解酶超家族中的一员,负责催化环裂产物C-C键的断裂,该反应是细菌降解芳香族化合物途径中的关键步骤. 为了解水解酶的催化特性,本文对该酶部分氨基酸进行了定点突变,并对突变体的动力学参数,化学修饰剂对突变体活性的影响以及突变体的二级结构进行了测定.各突变体的动力学参数特征为:突变体S110A,H265A和D237A的催化效率为野生型的1/104~1/103;突变体W85A和W219A催化效率分别为野生型的5/18和1/3,而同为色氨酸的突变体,W266A的催化效率只有野生型的1/104. 化学修饰剂对突变体S110A,H265A,D237A和W266A的酶活性几乎没有影响;而对突变体W85A和W219A却有较大的影响,修饰后,其相对活性仅为对照的10%~30%. 突变体的圆二色谱(CD谱)分析表明,与野生型相比,突变体的二级结构没有发生改变. 证明了Ser110,Asp237,His265是2-羟基-6-氧-6-苯基己-2,4-二烯酸水解酶(HOPDA hydrolase, HOPDA水解酶)催化反应所必需的氨基酸,并提出了Trp266在催化反应中也同样起到了非常关键的作用.  相似文献   

19.
C-C hydrolase MhpC (2-hydroxy-6-keto-nona-1,9-dioic acid 5,6-hydrolase) from Escherichia coli catalyses the hydrolytic C-C cleavage of the meta-ring fission product on the phenylpropionic acid catabolic pathway. The crystal structure of E. coli MhpC has revealed a number of active-site amino acid residues that may participate in catalysis. Site-directed mutants of His263, Ser110, His114, and Ser40 have been analysed using steady-state and stopped-flow kinetics. Mutants H263A, S110A and S110G show 10(4)-fold reduced catalytic efficiency, but still retain catalytic activity for C-C cleavage. Two distinct steps are observed by stopped-flow UV/Vis spectrophotometry, corresponding to ketonisation and C-C cleavage: H263A exhibits very slow ketonisation and C-C cleavage, whereas S110A and S110G exhibit fast ketonisation, an intermediate phase, and slow C-C cleavage. H114A shows only twofold-reduced catalytic efficiency, ruling out a catalytic role, but shows a fivefold-reduced K(M) for the natural substrate, and an ability to process an aryl-containing substrate, implying a role for His114 in positioning of the substrate. S40A shows only twofold-reduced catalytic efficiency, but shows a very fast (500 s(-1)) interconversion of dienol (317 nm) to dienolate (394 nm) forms of the substrate, indicating that the enzyme accepts the dienol form of the substrate. These data imply that His263 is responsible for both ketonisation of the substrate and for deprotonation of water for C-C cleavage, a novel catalytic role in a serine hydrolase. Ser110 has an important but non-essential role in catalysis, which appears not to be to act as a nucleophile. A catalytic mechanism is proposed involving stabilisation of reactive intermediates and activation of a nucleophilic water molecule by Ser110.  相似文献   

20.
The C1 domains of conventional and novel protein kinase C (PKC) isoforms bind diacylglycerol and phorbol esters with high affinity. Highly conserved hydrophobic residues at or near the rim of the binding cleft in the second cysteine-rich domain of PKC-delta (PKC-deltaC1b) were mutated to probe their roles in ligand recognition and lipid interaction. [(3)H]Phorbol 12,13-dibutyrate (PDBu) binding was carried out both in the presence and absence of phospholipids to determine the contribution of lipid association to the ligand affinity. Lipid dependence was determined as a function of lipid concentration and composition. The binding properties of a high affinity branched diacylglycerol with lipophilicity similar to PDBu were compared with those of PDBu to identify residues important for ligand selectivity. As expected, Leu-20 and Leu-24 strongly influenced binding. Substitution of either by aspartic acid abolished binding in either the presence or absence of phosphatidylserine. Mutation of Leu-20 to Arg or of Leu-24 to Lys caused a dramatic (340- and 250-fold, respectively) reduction in PDBu binding in the presence of lipid but only a modest reduction in the weaker binding of PDBu observed in the absence of lipid, suggesting that the main effect was on C1 domain -phospholipid interactions. Mutation of Leu-20 to Lys or of Trp-22 to Lys had modest (3-fold) effects and mutation of Phe-13 to Tyr or Lys was without effect. Binding of the branched diacylglycerol was less dependent on phospholipid and was more sensitive to mutation of Trp-22 to Tyr or Lys, especially in the presence of phospholipid, than was PDBu. In terms of specific PKC isoforms, our results suggest that the presence of Arg-20 in PKC-zeta may contribute to its lack of phorbol ester binding activity. More generally, the results emphasize the interplay between the C1 domain, ligand, and phospholipid in the ternary binding complex.  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