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 31 毫秒
1.
The crystal structures of cytochrome c peroxidase and ascorbate peroxidase are very similar, including the active site architecture. Both peroxidases have a tryptophan residue, designated the proximal Trp, located directly adjacent to the proximal histidine heme ligand. During the catalytic cycle, the proximal Trp in cytochrome c peroxidase is oxidized to a cation radical. However, in ascorbate peroxidase, the porphyrin is oxidized, not the proximal Trp, despite the close similarity between the two peroxidase active site structures. A cation located approximately 8 A from the proximal Trp in ascorbate peroxidase but absent in cytochrome c peroxidase is thought to be one reason why ascorbate peroxidase does not form a Trp radical. Site-directed mutagenesis has been used to introduce the ascorbate peroxidase cation binding site into cytochrome c peroxidase. Crystal structures show that mutants now bind a cation. Electron paramagnetic resonance spectroscopy shows that the cation-containing mutants of cytochrome c peroxidase no longer form a stable Trp radical. The activity of the cation mutants using ferrocytochrome c as a substrate is < 1% of wild type levels, while the activity toward a small molecule substrate, guaiacol, increases. These results demonstrate that long range electrostatic effects can control the reactivity of a redox active amino acid side chain and that oxidation/reduction of the proximal Trp is important in the oxidation of ferrocytochrome c.  相似文献   

2.
Two mutants of cytochrome c peroxidase (CCP) are reported which exhibit unique specificities toward oxidation of small substrates. Ala-147 in CCP is located near the delta-meso edge of the heme and along the solvent access channel through which H2O2 is thought to approach the active site. This residue was replaced with Met and Tyr to investigate the hypothesis that small molecule substrates are oxidized at the exposed delta-meso edge of the heme. X-ray crystallographic analyses confirm that the side chains of A147M and A147Y are positioned over the delta-meso heme position and might therefore modify small molecule access to the oxidized heme cofactor. Steady-state kinetic measurements show that cytochrome c oxidation is enhanced 3-fold for A147Y relative to wild type, while small molecule oxidation is altered to varying degrees depending on the substrate and mutant. For example, oxidation of phenols by A147Y is reduced to less than 20% relative to the wild-type enzyme, while Vmax/e for oxidation of other small molecules is less affected by either mutation. However, the "specificity" of aniline oxidation by A147M, i.e., (Vmax/e)/Km, is 43-fold higher than in wild-type enzyme, suggesting that a specific interaction for aniline has been introduced by the mutation. Stopped-flow kinetic data show that the restricted heme access in A147Y or A147M slows the reaction between the enzyme and H202, but not to an extent that it becomes rate limiting for the oxidation of the substrates examined. The rate constant for compound ES formation with A147Y is 2.5 times slower than wild-type CCP. These observations strongly support the suggestion that small molecule oxidations occur at sites on the enzyme distinct from those utilized by cytochrome c and that the specificity of small molecule oxidation can be significantly modulated by manipulating access to the heme edge. The results help to define the role of alternative electron transfer pathways in cytochrome c peroxidase and may have useful applications in improving the specificity of peroxidase with engineered function.  相似文献   

3.
Presteady and steady-state kinetic results on the interactions of a wild-type, and the mutant glucoamylases Trp52-->Phe and Trp317-->Phe, from Aspergillus niger with maltose, maltotriose and maltotetraose have been obtained and analyzed. The results are compared with previous ones on the mutants, Trp120-->Phe and Glu180-->Gln, and with results obtained from structure energy minimization calculations based on known three-dimensional structural data. All results are in accordance with a three-step reaction model involving two steps in the substrate binding and a rate-determining catalytic step. Trp317 and Glu180 belong to different subsites, but are placed on the same flank of the active site (beta-flank). The Trp317-->Phe and the Glu180-->Gln mutants show almost identical kinetic results: weakening of the substrate binding, mainly caused by changes in the second reaction step, and practically no change of the catalytic rate. Structure energy minimization calculations show that the same loss of Arg305 and Glu180 hydrogen bonds to the substrate occurs in the Michaelis complexes of each of these mutants. These results indicate that important interactions of the active site may be better understood from a consideration of its flanks rather than of its subsites. The results further indicate differences in the substrate binding mode of maltose and of longer substrates. Trp52 and Trp120 each interact with the catalytic acid, Glu179, and are placed on the flank (alpha-flank) of the active site opposite to Trp317, Arg305 and Glu180. Also the Trp52-->Phe and Trp120-->Phe mutants show kinetic results similar to each other. The catalytic rates are strongly reduced and the substrates are bound more strongly, mainly as a result of the formation of a more stable complex in the second reaction step. All together, the substrate binding mechanism seems to involve an initial enzyme-substrate complex, in which the beta-flank plays a minor role, except for maltose binding; this is followed by a conformational change, in which hydrogen bonds to Arg305 and Glu180 of the beta-flank are established and the correct alignment on the alpha-flank of Glu179, the general acid catalyst, governed by its flexible interactions with Trp52 and Trp120, occurs.  相似文献   

4.
In the high-resolution crystal structures of two lignin peroxidase isozymes from the white rot fungus Phanerochaete chrysosporium a significant electron density at single bond distance from the C beta of Trp171 was observed and interpreted as a hydroxy group. To further clarify the nature of this feature, we carried out tryptic digestion of the enzyme and isolated the Trp171 containing peptide. Under ambient conditions, this peptide shows an absorbance spectrum typical of tryptophan. At elevated temperature, however, the formation of an unusual absorbance spectrum with lambda max = 333 nm can be followed that is identical to that of N-acetyl-alpha, beta-didehydrotryptophanamide, resulting upon water elimination from beta-hydroxy tryptophan. The Trp171 containing tryptic peptide isolated from the recombinant and refolded lignin peroxidase produced from Escherichia coli does not contain the characteristic 333 nm absorbance band at any temperature. However, treatment with 3 equiv of H2O2 leads to complete hydroxylation of Trp171. Reducing substrates compete with this process, e.g., in the presence of 0.5 mM veratryl alcohol, about 7 equiv of H2O2 is necessary for complete modification. We conclude that the hydroxylation at the C beta of Trp171 is an autocatalytic reaction which occurs readily under conditions of natural turnover, e.g., in the ligninolytic cultures of P. chrysosporium, which are known to contain an oxidase-based H2O2-generating system. No dependence on dioxygen was found for this oxidative process. Chemical modification of fungal lignin peroxidase with the tryptophan-specific agent N-bromo succinimide leads to a drastically reduced activity with respect to the substrate veratryl alcohol. This suggests that Trp171 is involved in catalysis and that electron transfer from this surface residue to the oxidized heme cofactor is possible under steady-state conditions.  相似文献   

5.
Site-directed mutagenesis was used to identify the veratryl alcohol binding site of lignin peroxidase. The cDNA encoding isozyme H8 was mutated at Glu146 to both an Ala and a Ser residue. The H8 polypeptide was produced by E. coli as inclusion bodies and refolded to yield active enzyme. The wild type recombinant enzyme and the mutants were purified to homogeneity and characterized by steady state kinetics. The kcat is decreased for both mutants of Glu146. The reactivity of mutants (kcat/Km) toward H2O2 were not affected. In contrast, the kcat/Km of the mutants for veratryl alcohol were decreased by at least half. The oxidation of guaiacol by these mutants were more significantly affected. These results collectively suggest that E146 plays a central role in the binding of veratryl alcohol by lignin peroxidase.  相似文献   

6.
Previous thermodynamic and computational studies have pointed to the important energetic role of aromatic contacts in generating the exceptional binding free energy of streptavidin-biotin association. We report here the crystallographic characterization of single site tryptophan mutants in investigating structural consequences of alterations in these aromatic contacts. Four tryptophan residues, Trp79, Trp92, Trp108 and Trp120, play an important role in the hydrophobic binding contributions, which along with a hydrogen bonding network and a flexible binding loop give rise to tight ligand binding (Ka approximately 10(13) M-1). The crystal structures of ligand-free and biotin-bound mutants, W79F, W108F, W120F and W120A, in the resolution range from 1.9 to 2.3 A were determined. Nine data sets for these four different mutants were collected, and structural models were refined to R-values ranging from 0.15 to 0.20. The major question addressed here is how these mutations influence the streptavidin binding site and in particular how they affect the binding mode of biotin in the complex. The overall folding of streptavidin was not significantly altered in any of the tryptophan mutants. With one exception, only minor deviations in the unbound structures were observed. In one crystal form of unbound W79F, there is a coupled shift in the side-chains of Phe29 and Tyr43 toward the mutation site, although in a different crystal form these shifts are not observed. In the bound structures, the orientation of biotin in the binding pocket was not significantly altered in the mutant complex. Compared with the wild-type streptavidin-biotin complex, there were no additional crystallographic water molecules observed for any of the mutants in the binding pocket. These structural studies thus suggest that the thermodynamic alterations can be attributed to the local alterations in binding residue composition, rather than a rearrangement of binding site architectures.  相似文献   

7.
The role of amino acid residues located in the active site pocket of phosphatidylinositol-specific phospholipase C (PI-PLC) from Bacillus cereus[Heinz, D. W., Ryan, M., Bullock, T., & Griffith, O. H. (1995) EMBO J. 14, 3855-3863] was investigated by site-directed mutagenesis, kinetics, and crystal structure analysis. Twelve residues involved in catalysis and substrate binding (His32, Arg69, His82, Gly83, Lys115, Glu117, Arg163, Trp178, Asp180, Asp198, Tyr200, and Asp274) were individually replaced by 1-3 other amino acids, resulting in a total number of 21 mutants. Replacements in the mutants H32A, H32L, R69A, R69E, R69K, H82A, H82L, E117K, R163I, D198A, D198E, D198S, Y200S, and D274S caused essentially complete inactivation of the enzyme. The remaining mutants (G83S, K115E, R163K, W178Y, D180S, Y200F, and D274N) exhibited reduced activities up to 57% when compared with wild-type PI-PLC. Crystal structures determined at a resolution ranging from 2.0 to 2.7 A for six mutants (H32A, H32L, R163K, D198E, D274N, and D274S) showed that significant changes were confined to the site of the respective mutation without perturbation of the rest of the structure. Only in mutant D198E do the side chains of two neighboring arginine residues move across the inositol binding pocket toward the newly introduced glutamic acid. An analysis of these structure-function relationships provides new insight into the catalytic mechanism, and suggests a molecular explanation of some of the substrate stereospecificity and inhibitor binding data available for this enzyme.  相似文献   

8.
Mutational analysis of human uroporphyrinogen decarboxylase   总被引:1,自引:0,他引:1  
Uroporphyrinogen decarboxylase (URO-D), a heme biosynthetic enzyme, catalyzes the multi-step decarboxylation reaction converting uroporphyrinogen I or III to coproporphyrinogen I or III. The URO-D protein has been purified from several sources and its gene has been cloned from many organisms. In spite of this, little is known about the active site(s) of the enzyme. Inhibitor studies suggest that cysteine and histidine residues are important for enzyme activity. We employed the Kunkel method of site-directed mutagenesis to convert each of the six cysteines in human URO-D to serine and each of the three conserved histidines to asparagine. Recombinant mutant URO-D's were expressed in Escherichia coli, partially purified, and their kinetic properties compared to recombinant wild-type URO-D. All cysteine mutants retained approx. 40% wild-type enzyme activity, indicating that no single cysteine is absolutely critical for the integrity of the catalytic site. The three histidine mutants also retained significant enzyme activity and one, (H339N), displayed unique properties. The H339N mutation resulted in an enzyme with high residual activity but decarboxylation of intermediate reaction products of the I isomer series was markedly abnormal. The histidine at residue 339 is likely important in imparting isomer specificity.  相似文献   

9.
Site-directed mutagenesis was used to experimentally verify the 3D model of the active site of herpes simplex virus type-1 thymidine kinase (HSV 1 TK) obtained by homology modelling. For this purpose, D215 and K317 were replaced by R and G, respectively, at homologous positions in the aciclovir-insensitive bovine herpes virus type-1 thymidine kinase (BHV 1 TK). Wild-type and mutated enzymes were expressed in Escherichia coli using a gene fusion vector and purified to homogeneity. While both mutants had the same Km value for thymidine as the recombinant wild-type enzyme (0.2 microM), Vmax was decreased to 20-25% of the original wild-type value. The recombinant wild-type enzyme was inhibited by the substrate analogue aciclovir with a Ki of 146 microM. Both mutants were able to phosphorylate aciclovir to about the same extent as the wild-type enzyme. These findings suggest that neither D215 nor K317 are directly involved in substrate binding. Therefore, a rearrangement of the 3D model is suggested, concerning the assignment of the substrate-binding site and co-substrate-binding site at the right and left side of the phosphate-binding loop, respectively.  相似文献   

10.
The molecular mechanism of substrate recognition in membrane transport is not well understood. Two amino acid residues, Tyr446 and Trp455 in transmembrane segment 10 (TM10), have been shown to be important for galactose recognition by the yeast Gal2 transporter; Tyr446 was found to be essential in that its replacement by any of the other 19 amino acids abolished transport activity (Kasahara, M., Shimoda, E., and Maeda, M. (1997) J. Biol. Chem. 272, 16721-16724). The Glut1 glucose transporter of animal cells belongs to the same Glut transporter family as does Gal2 and thus might be expected to show a similar mechanism of substrate recognition. The role of the two amino acids, Phe379 and Trp388, in rat Glut1 corresponding to Tyr446 and Trp455 of Gal2 was therefore studied. Phe379 and Trp388 were individually replaced with each of the other 19 amino acids, and the mutant Glut1 transporters were expressed in yeast. The expression level of most mutants was similar to that of the wild-type Glut1, as revealed by immunoblot analysis. Glucose transport activity was assessed by reconstituting a crude membrane fraction of the yeast cells in liposomes. No significant glucose transport activity was observed with any of Trp388 mutants, whereas the Phe379 mutants showed reduced or no activity. These results indicate that the two aromatic amino acids in TM10 of Glut1 are important for glucose transport. However, unlike Gal2, the residue at the cytoplasmic end of TM10 (Trp388, corresponding to Trp455 of Gal2), rather than that in the middle of TM10 (Phe379, corresponding to Tyr446 of Gal2), is essential for transport activity.  相似文献   

11.
In the accompanying paper, we demonstrated the presence of a fluorescence resonance energy transfer (FRET) between the tryptophans of the melibiose permease (MelB) of Escherichia coli and a fluorescent sugar, 2'-(N-5-dimethylaminonaphthalene-1-sulfonyl)aminoethyl-1-thio-beta-D- galactopyranoside (Dns2-S-Gal) bound at the sugar-binding site (Maehrel, C., Cordat, E., Mus-Veteau, I., and Leblanc, G. (1998) J. Biol. Chem. 273, 33192-33197). To identify the tryptophans that transfer their energy to the fluorescent sugar, we analyzed the FRET properties of MelB mutants carrying the replacement of each of the eight MelB tryptophans by a phenylalanine. The data indicate that Trp64, localized in loop 2-3 from the N-terminal domain, and Trp299, localized in helix IX in the C-terminal domain, are responsible for up to 80% of the FRET signal. Moreover, by assuming that only Trp299 transfers energy to Dns2-S-Gal in mutant W64F, whereas only Trp64 transfers energy to Dns2-S-Gal in mutant W299F, we calculated that Trp299 and Trp64 are about 14 and 20 A away from the probe, respectively. In addition, we observed that mutating Trp342, localized in helix X of the C-terminal domain, produces a significant increase of the polarity of the fluorescent sugar environment, suggesting its proximity to the sugar-binding site. Taken together, these data provide additional support for the suggestion that (i) the sugar-binding site is localized in the C-terminal part of the transporter, probably close to membrane segments IX and X, and (ii) the N-terminal domain, and particularly cytoplasmic loop 2-3, is also close to the sugar-binding site.  相似文献   

12.
A haem peroxidase different from other microbial, plant and animal peroxidases is described. The enzyme is secreted as two isoforms by dikaryotic Pleurotus eryngii in peptone-containing liquid medium. The corresponding gene, which presents 15 introns and encodes a 361-amino-acid protein with a 30-amino-acid signal peptide, was isolated as two alleles corresponding to the two isoforms. The alleles differ in three amino acid residues and in a seven nucleotide deletion affecting a single metal response element in the promoter. When compared with Phanerochaete chrysosporium peroxidases, the new enzyme appears closer to lignin peroxidase (LiP) than to Mn-dependent peroxidase (MnP) isoenzymes (58-60% and 55% identity respectively). The molecular model built using crystal structures of three fungal peroxidases as templates, also showed high structural affinity with LiP (C alpha-distance 1.2 A). However, this peroxidase includes a Mn2+ binding site formed by three acidic residues (E36, E40 and D175) near the haem internal propionate, which accounts for the ability to oxidize Mn2+. Its capability to oxidize aromatic substrates could involve interactions with aromatic residues at the edge of the haem channel. Another possibility is long-range electron transfer, e.g. from W164, which occupies the same position of LiP W171 recently reported as involved in the catalytic cycle of LiP.  相似文献   

13.
Cytochrome bo is a member of the heme-copper terminal oxidase superfamily and serves as a four-subunit ubiquinol oxidase in the aerobic respiratory chain of Escherichia coli. To probe the location and structural properties of the ubiquinol oxidation site, we isolated and characterized five or 10 spontaneous mutants resistant to either 2,6-dimethyl-1,4-benzoquinone, 2,6-dichloro-4-nitrophenol, or 2,6-dichloro-4-dicyanovinylphenol, the potent competitive inhibitors for the oxidation of ubiquinol-1 [Sato-Watanabe, M., Mogi, T., Miyoshi, H., Iwamura, H., Matsushita, K., Adachi, O., and Anraku, Y. (1994) J. Biol. Chem. 269, 28899-28907]. Analyses of the growth yields and the ubiquinol-1 oxidase activities of the mutant membranes showed that the mutations increased the degree of the resistance to the selecting compounds. Notably, several mutants showed the cross-resistance. These data indicate that the binding sites for substrate and the competitive inhibitors are partially overlapped in the ubiquinol oxidation site. All the mutations were linked to the expression vector, and 23 mutations examined were all present in the C-terminal hydrophilic domain (Pro96-His315) of subunit II. Sequencing analysis revealed that seven mutations examined are localized near both ends of the cupredoxin fold. Met248Ile, Ser258Asn, Phe281Ser, and His284Pro are present in a quinol oxidase-specific (Qox) domain and proximal to low-spin heme b in subunit I and the lost CuA site in subunit II, whereas Ile129Thr, Asn198Thr, and Gln233His are rather scattered in a three-dimensional structure and closer to transmembrane helices of subunit II. Our data suggest that the Qox domain and the CuA end of the cupredoxin fold provide the quinol oxidation site and are involved in electron transfer to the metal centers in subunit I.  相似文献   

14.
The crucial reaction intermediate in the reaction of peroxidase with hydrogen peroxide (H2O2), compound I, contains a porphyrin pi-cation radical in horseradish peroxidase (HRP), which catalyzes oxidation of small organic and inorganic compounds, whereas cytochrome c peroxidase (CcP) has a radical center on the tryptophan residue (Trp-191) and oxidizes the redox partner, cytochrome c. To investigate the roles of the amino acid residue near the heme active center in discriminating the function of the peroxidases in these two enzymes, we prepared a CcP-like HRP mutant, F221W (Phe-221 --> Trp). Although the rapid spectral scanning and stopped-flow experiments confirmed that the F221W mutant reacts with H2O2 to form the porphyrin pi-cation radical at the same rate as for the wild-type enzyme, the characteristic spectral features of the porphyrin pi-cation radical disappeared rapidly, and were converted to the compound II-type spectrum. The EPR spectrum of the resultant species produced by reduction of the porphyrin pi-cation radical, however, was quite different from that of compound II in HRP, showing typical signals from a Trp radical as found for CcP. The sequential radical formation from the porphyrin ring to the Trp residue implies that the proximal Trp is a key residue in the process of the radical transfer from the porphyrin ring, which differentiates the function of peroxidases.  相似文献   

15.
The presence of a two-subunit cytochrome (cyt) b-c1 subcomplex in chromatophore membranes of Rhodobacter capsulatus mutants lacking the Rieske iron-sulfur (Fe-S) protein has been described previously [Davidson, E., Ohnishi, T., Tokito, M., and Daldal, F. (1992) Biochemistry 31, 3351-3358]. Here, this subcomplex was purified to homogeneity in large quantities, and its properties were characterized. As expected, it contained stoichiometric amounts of cyt b and cyt c1 subunits forming a stable entity devoid of the Fe-S protein subunit. The spectral and thermodynamic properties of its heme groups were largely similar to those of a wild-type bc1 complex, except that those of its cyt bL heme were modified as revealed by EPR spectroscopy. Dark potentiometric titrations indicated that the redox midpoint potential (Em7) values of cytochromes bH, bL, and c1 were very similar to those of a wild-type bc1 complex. The purified b-c1 subcomplex had a nonfunctional ubihydroquinone (UQH2) oxidation (Qo) site, but it contained an intact ubiquinone (UQ) reductase (Qi) site as judged by its ability to bind the Qi inhibitor antimycin A, and by the presence of antimycin A sensitive Qi semiquinone. Interestingly, its Qo site could be readily reconstituted by addition of purified Fe-S protein subunit. Reactivated complex exhibited myxothiazol, stigmatellin, and antimycin A sensitive cyt c reductase activity and an EPR gx signal comparable to that observed with a bc1 complex when the Qo site is partially occupied with UQ/UQH2. However, a mutant derivative of the Fe-S protein subunit lacking its first 43 amino acid residues was unable to reactivate the purified b-c1 subcomplex although it could bind to its Qo site in the presence of stigmatellin. These findings demonstrated for the first time that the amino-terminal membrane-anchoring domain of the Fe-S protein subunit is necessary for UQH2 oxidation even though its carboxyl-terminal domain is sufficient to provide wild-type-like interactions with stigmatellin at the Qo site of the bc1 complex.  相似文献   

16.
Six single-Trp mutants were engineered by individually reintroducing each of the native Trp residues into a functional lactose permease mutant devoid of Trp (Trp-less permease; Menezes ME, Roepe PD, Kaback HR, 1990, Proc Natl Acad Sci USA 87:1638-1642), and fluorescent properties were studied with respect to solvent accessibility, as well as alterations produced by ligand binding. The emission of Trp 33, Trp 78, Trp 171, and Trp 233 is strongly quenched by both acrylamide and iodide, whereas Trp 151 and Trp 10 display a decrease in fluorescence in the presence of acrylamide only and no quenching by iodide. Of the six single-Trp mutants, only Trp 33 exhibits a significant change in fluorescence (ca. 30% enhancement) in the presence of the substrate analog beta,D-galactopyranosyl 1-thio-beta,D-galactopyranoside (TDG). This effect was further characterized by site-directed fluorescent studies with purified single-Cys W33-->C permease labeled with 2-(4'-maleimidylanilino)-naphthalene-6-sulfonic acid (MIANS). Titration of the change in the fluorescence spectrum reveals a 30% enhancement accompanied with a 5-nm blue shift in the emission maximum, and single exponential behavior with an apparent KD of 71 microM. The effect of substrate binding on the rate of MIANS labeling of single-Cys 33 permease was measured in addition to iodide and acrylamide quenching of the MIANS-labeled protein. Complete blockade of labeling is observed in the presence of TDG, as well as a 30% decrease in accessibility to iodide with no change in acrylamide quenching. Overall, the findings are consistent with the proposal (Wu J, Frillingos S, Kaback HR, 1995a, Biochemistry 34:8257-8263) that ligand binding induces a conformational change at the C-terminus of helix I such that Pro 28 and Pro 31, which are on one face, become more accessible to solvent, whereas Trp 33, which is on the opposite face, becomes less accessible to the aqueous phase. The findings regarding accessibility to collisional quenchers are also consistent with the predicted topology of the six native Trp residues in the permease.  相似文献   

17.
Single OmpC porin channels have been reconstituted in planar bilayer membranes. Wild-type OmpC forms trimers which are largely insensitive to voltages below 250 mV.A single-point mutation of the ompC gene has been prepared resulting in replacement of Trp56 by Cys in the eyelet region of the channel wall in a highly conserved segment of the polypeptide. The monomeric channels of which the trimer is composed have smaller conductivity in 1 M NaCl (400 +/- 20 pS, mean and S.E.M., n=30) and increased voltage sensitivity by comparison with the wild-type under similar conditions, whereas other (Dex) mutants form larger channels and display different behaviour. Further, by treatment in SDS solutions at different temperatures, the W56C mutant has been shown to be less stable than either the wild-type or the Dex mutants.  相似文献   

18.
5-Aminolevulinate synthase (EC 2.3.1.37) catalyzes the first reaction in the heme biosynthetic pathway in nonplant eukaryotes and some prokaryotes. Homology sequence modeling between 5-aminolevulinate synthase and some other alpha-family pyridoxal 5'-phosphate-dependent enzymes indicated that the residue corresponding to the Arg-439 of murine erythroid 5-aminolevulinate synthase is a conserved residue in this family of pyridoxal 5'-phosphate-dependent enzymes. Further, this conserved arginine residue in several enzymes, e.g., aspartate aminotransferase, for which the three-dimensional structure is known, has been shown to interact with the substrate carboxyl group. To test whether Arg-439 is involved in substrate binding in murine erythroid 5-aminolevulinate synthase, Arg-439 and Arg-433 of murine erythroid 5-aminolevulinate synthase were each replaced by Lys and Leu using site-directed mutagenesis. The R439K mutant retained 77% of the wild-type activity; its K(m) values for both substrates increased 9-13-fold, while the activity of R433K increased 2-fold and the K(m) values for both substrates remained unchanged. R439L had no measurable activity as determined using a standard 5-aminolevulinate synthase enzyme-coupled activity assay. In contrast, the kinetic parameters for R433L were comparable to those of the wild-type. Dissociation constants (Kd) for glycine increased 5-fold for R439K and at least 30-fold for R439L, while Kd values for glycine for both R433K and R433L mutants were similar to those of the wild-type. However, there was not much difference in methylamine binding among the mutants and the wild-type, excepting of a 10-fold increase in K(d)methylamine for R439L. R439K proved much less thermostable than the wild-type enzyme, with the thermotransition temperature, T1/2, determined to be 8.3 degrees C lower than that of the wild-type enzyme. In addition, in vivo complementation analysis demonstrated that in the active site of murine erythroid 5-aminolevulinate synthase, R439 is contributed from the same subunit as K313 (which is involved in the Schiff base linkage of the pyridoxal 5'-phosphate cofactor) and D279 (which interacts electrostatically with the ring nitrogen of the cofactor), while another subunit provides R149. Taken together, these findings suggest that Arg-439 plays an important role in substrate binding of murine erythroid 5-aminolevulinate synthase.  相似文献   

19.
A number of thrombin mutants have been constructed to investigate the role of Trp96 and the beta-insertion loop for the specificity of thrombin. Thrombin(60D) consists of the replacement of the beta-insertion loop (14 amino acid residues from 59 to 63, including a 9-residue insertion at position 60) with the corresponding four residues in trypsin, Tyr-Lys-Ser-Gly; thrombin(GGG) is a smaller loop mutation in which the residues Tyr(60A)Pro(60B)Pro(60C)Trp(60D) Asp(60E)Lys(60F) of the beta-insertion loop were replaced by Gly-Gly-Gly; thrombin(96S) consists of a point mutation Trp96 --> Ser; and thrombin(GGG/96S) is the double mutant incorporating both changes. Thrombin(96S) clots fibrinogen approximately 3 times more slowly than thrombin, with the two beta-insertion loop mutants, thrombin(GGG) and thrombin(GGG/96S), reacting approximately 3000- and 1300-fold more slowly, respectively. The specificity constant kcat/Km for the cleavage of fibrinopeptide A and fibrinopeptide B by thrombin(96S) was 2.6 and 0.35 microM(-1) s(-1), respectively, compared to 10 and 2.5 microM(-1) s(-1) for wild-type recombinant thrombin, respectively. Kinetic constants were determined for the hydrolysis of H-D-phenylalanyl-L-pipecolyl-L-arginine-p-nitroaniline. The Michaelis constant Km increased approximately 6-fold for thrombin(96S) and >200-fold for thrombin(GGG) and thrombin(GGG/96S) when compared to wild-type recombinant thrombin, while the catalytic constant kcat remained approximately the same. All mutants were more susceptible to inhibition by BPTI than wild-type recombinant thrombin. Clearly, the beta-insertion loop is important for thrombin activity. But the mutation of Trp96 --> Ser can compensate somewhat for the loss of binding at the beta-insertion loop. The deletion of the hydrophobic interaction between Trp96 and Pro(60B)Pro(60C) appears to decrease the stability of the beta-insertion loop, thereby causing a decrease in binding efficiency.  相似文献   

20.
The phosphotriesterase from Pseudomonas diminuta hydrolyzes a wide variety of organophosphate insecticides and acetylcholinesterase inhibitors. The rate of hydrolysis depends on the substrate and can range from 6000 s-1 for paraoxon to 0.03 s-1 for the slower substrates such as diethylphenylphosphate. Increases in the reactivity of phosphotriesterase toward the slower substrates were attempted by the placement of a potential proton donor group at the active site. Distances from active site residues in the wild type protein to a bound substrate analog were measured, and Trp131, Phe132, and Phe306 were found to be located within 5.0 A of the oxygen atom of the leaving group. Eleven mutants were created using site-directed mutagenesis and purified to homogeneity. Phe132 and Phe306 were replaced by tyrosine and/or histidine to generate all combinations of single and double mutants at these two sites. The single mutants W131K, F306K, and F306E were also constructed. Kinetic constants were measured for all of the mutants with the substrates paraoxon, diethylphenylphosphate, acephate, and diisopropylfluorophosphate. Vmax values for the mutant enzymes with the substrate paraoxon varied from near wild type values to a 4-order of magnitude decrease for the W131K mutant. There were significant increases in the Km for paraoxon for all mutants except F132H. Vmax values measured using diethylphenylphosphate decreased for all mutants except for F132H and F132Y, whereas Km values ranged from near wild type levels to increases of 25-fold. Vmax values for acephate hydrolysis ranged from near wild type values to a 10(3)-fold decrease for W131K. Km values for acephate ranged from near wild type to a 5-fold increase. Vmax values for the mutants tested with the substrate diisopropylfluorophosphate showed an increase in all cases except for the W131K, F306K, and F306E mutants. The Vmax value for the F132H/F306H mutant was increased to 3100 s-1. These studies demonstrated for the first time that it is possible to significantly enhance the ability of the native phosphotriesterase to hydrolyze phosphorus-fluorine bonds at rates that rival the hydrolysis of paraoxon.  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