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.
Utilizing site-directed mutagenesis, 77 charged and polar residues that are highly exposed on the surface of human thrombin were systematically substituted with alanine. Functional assays using thrombin mutants identified residues that were required for the recognition and cleavage of the procoagulant substrate fibrinogen (Lys21, Trp50, Lys52, Asn53 + Thr55, Lys65, His66, Arg68, Tyr71, Arg73, Lys77, Lys106 + Lys107, Asp193 + Lys196, Glu202, Glu229, Arg233, Asp234) and the anticoagulant substrate protein C (Lys21, Trp50, Lys65, His66, Arg68, Tyr71, Arg73, Lys77, Lys106 + Lys107, Glu229, Arg233), interactions with the cofactor thrombomodulin (Gln24, Arg70) and inhibition by the thrombin aptamer, an oligonucleotide-based thrombin inhibitor (Lys65, His66, Arg70, Tyr71, Arg73). Although there is considerable overlap between the functional epitopes, distinct and specific residues with unique functions were identified. When the functional residues were mapped on the surface of thrombin, they were located on a single hemisphere of thrombin that included both the active site cleft and the highly basic exosite 1. No functional residues were located on the opposite face of thrombin. Residues with procoagulant or anticoagulant functions were not spatially separated but interdigitated with residues of opposite or shared function. Thus thrombin utilizes the same general surface for substrate recognition regardless of substrate function although the critical contact residues may vary.  相似文献   

2.
The flavonol 3- and 4'-sulfotransferases (ST) from Flaveria chloraefolia catalyze the transfer of the sulfonate group from 3'-phosphoadenosine 5'-phosphosulfate (PAdoPS) to position 3 of flavonol aglycones and position 4' of flavonol 3-sulfates. We identified previously a protein segment, designated domain II, that contains all the determinants responsible for the specificity of these enzymes. Within domain II, at least five amino acids specific to the 4'-ST that could bind the sulfate group of quercetin 3-sulfate were identified. In this study, these amino acid residues were introduced at equivalent positions in the flavonol 3-ST sequence by site-directed mutagenesis of the cloned cDNA. No reversal of the substrate specificity was observed after the individual mutations. However, mutation of Leu95 to Tyr had different effects on the kinetic constants depending on the substitution pattern of the flavonoid B ring, suggesting that the tyrosine side chain may be in direct contact with this part of the molecule. The function of conserved amino acids present in domain II was also investigated. Unconservative mutations at Lys134, Tyr137 and Tyr150 resulted in protein instability in solution, suggesting that these residues might be important for the structural stability of the enzyme. Replacement of Arg140 with Lys or Ser had no effect on protein stability, but resulted in a strong reduction in specific activity. The results of photoaffinity-labeling experiments with PAdoP[35S]S suggest that this residue is required to bind the cosubstrate. In addition, the reduced affinity of [Ser140]ST for 3'-phosphoadenosine 5'-phosphate (PAdoP)-agarose indicates that Arg140 is also involved in binding the coproduct. Replacement of His118 with Glu or Ala resulted in a strong reduction in catalytic activity. However, [Lys118]ST retained a significant amount of catalytic activity. The results of photoaffinity-labeling experiments with PAdoP[35S]S and affinity chromatography on PAdoP-agarose suggest that His118 might be involved in catalysis in the flavonol 3-ST.  相似文献   

3.
The mechanism of autooxidation of myoglobin   总被引:1,自引:0,他引:1  
Time courses for the autooxidation of native and mutant sperm whale and pig myoglobins were measured at 37 degrees C in the presence of catalase and superoxide dismutase. In sperm whale myoglobin, His64(E7) was replaced with Gln, Gly, Ala, Val, Thr, Leu, and Phe; Val68(E11) was replaced with Ala, Ile, Leu, and Phe; Leu29(B10) was replaced with Ala, Val, and Phe. In pig myoglobin, His64(E7) was replaced with Val; Val68(E11) was replaced with Thr and Ser; Thr67(E10) was replaced with Ala, Val, Glu, and Arg; Lys45(CD3) was replaced with Ser, Glu, His, and Arg. The observed pseudo-first order rate constants varied over 4 orders of magnitude, from 58 h-1 (H64A) to 0.055 h-1 (native) to 0.005 h-1 (L29F) at 37 degrees C, pH 7, in air. The dependences of the observed autooxidation rate constant on oxygen concentration and pH were measured for native and selected mutant myoglobins. In the native proteins and in most mutants still possessing the distal histidine, autooxidation occurs through a combination of two mechanisms. At high [O2], direct dissociation of the neutral superoxide radical (HO2) from oxymyoglobin dominates, and this process is accelerated by decreasing pH. At low [O2], autooxidation occurs by a bimolecular reaction between molecular oxygen and deoxymyoglobin containing a weakly coordinated water molecule. The neutral side chain of the distal histidine (His64) inhibits autooxidation by hydrogen bonding to bound oxygen, preventing both HO2 dissociation and the oxidative bimolecular reaction with deoxymyoglobin. Replacement of His64 by amino acids incapable of hydrogen bonding to the bound ligand markedly increases the rate of autooxidation and causes the superoxide mechanism to predominate. Increasing the polarity of the distal pocket by substitution of Val68 with Ser and Thr accelerates autooxidation, presumably by facilitating protonation of the Fe(II).O2 complex. Increasing the net anionic charge at the protein surface in the vicinity of the heme group also enhances the rate of autooxidation. Decreasing the volume of the distal pocket by replacing small amino acids with larger aliphatic or aromatic residues at positions 68 (E11) and 29 (B10) inhibits autooxidation markedly by decreasing the accessibility of the iron atom to solvent water molecules.  相似文献   

4.
Mice constitutively express glutathione S-transferase mGSTA3-3 in liver. This isoform possesses uniquely high conjugating activity toward aflatoxin B1-8,9-epoxide (AFBO), thereby protecting mice from aflatoxin B1-induced hepatocarcinogenicity. In contrast, rats constitutively express a closely related GST isoenzyme, rGSTA3-3, with low AFBO activity and, therefore, are sensitive to aflatoxin B1 exposure. Although the two GSTs share 86% sequence identity and have similar catalytic activities toward 1-chloro-2,4-dinitrobenzene (CDNB), they have an approximately 1000-fold difference in catalytic activity toward AFBO. To identify amino acids that confer high activity toward AFBO, non-conserved rGSTA3-3 residues were replaced with mGSTA3-3 residues in two regions believed to form the substrate binding site. Twenty-one mutant rGSTA3-3 enzymes were generated by site-directed mutagenesis using combinations of nine different residues. Except for the E208D mutant, single mutations of rGSTA3-3 produced enzymes with no detectable AFBO activity. Generally, AFBO conjugation activity increased in additive fashion as mGSTA3-3 residues were introduced into the rGSTA3-3 enzyme with the six site mutant E104I/H108Y/Y111H/L207F/E208D/V217K displaying the highest AFBO activity (40 nmol/mg/min) of all the mutant enzymes. When this mutant enzyme was further modified by three additional substitutions (D103E/I105M/V106I) AFBO conjugation activity decreased 14-fold to 2. 8 nmol/mg/min. Although wild-type mGSTA3-3 AFBO conjugation activity (265 nmol/mg/min) could not be obtained by our rGSTA3-3 mutants, we were able to identify six mGSTA3-3 residues; Ile104, Tyr108, His111, Phe207, Asp208, and Lys217 that, when collectively substituted into rGSTA3-3, substantially increased (>200-fold) glutathione conjugation activity toward AFBO.  相似文献   

5.
Previous studies have demonstrated that the in vitro folding pathway of Escherichia coli serine hydroxymethyltransferase has both monomer and dimer intermediates that are stable for periods of minutes to hours at 4 degrees C (Cai K., Schirch, D., and Schirch, V. (1995) J. Biol. Chem. 270, 19294-19299). Single Trp mutant enzymes were constructed and used in combination with other methods to show that on the folding pathway of this enzyme two domains rapidly fold to form a monomer in which the amino-terminal 55 amino acid residues and a segment around the active site region of Lys229 remain in a largely disordered form. This partially folded enzyme can form dimers and slowly undergoes a rate-determining conformational change in which the unstructured segments assume their native state (Cai, K. , and Schirch, V. (1996) J. Biol. Chem. 271, 2987-2994). To further assess the kinetics and structural details of the intermediates during folding, fluorescence energy transfer and fluorescence anisotropy measurements were made of the three Trp residues and pyridoxal 5'-phosphate, attached covalently to the active site by reduction to a secondary amine by sodium cyanoborohydride. These studies confirmed that the basic kinetic folding pathway remained the same in the reduced enzyme as compared to the earlier studies with the apoenzyme. Both equilibrium and kinetic intermediates were identified and their structural characteristics determined. The results show that the active site Lys229-bound pyridoxyl 5'-phosphate remains more than 50 angstroms from any Trp residues until the final rate-determining conformational change when it approaches each Trp residue at the same rate. The environment of each Trp residue and the pyridoxyl phosphate in both an equilibrium folding intermediate and a kinetic folding intermediate are described.  相似文献   

6.
The active-site cleft of bovine pancreatic ribonuclease A (RNase A) is lined with cationic residues that interact with a bound nucleic acid. Those residues interacting with the phosphoryl groups comprise the P0, P1, and P2 subsites, with the scissile P-O5' bond residing in the P1 subsite. Coulombic interactions between the P0 and P2 subsites and phosphoryl groups of the substrate were characterized previously [Fisher, B. M., Ha, J.-H., and Raines, R. T. (1998) Biochemistry 37, 12121-12132]. Here, the interactions between these subsites and the active-site residues His12 and His119 are described in detail. A protein variant in which the cationic residues in these subsites (Lys66 in the P0 subsite and Lys7 and Arg10 in the P2 subsite) were replaced with alanine was crystallized, both free and with bound 3'-uridine monophosphate (3'-UMP). Structures of K7A/R10A/K66A RNase A and the K7A/R10A/K66A RNase A.3'-UMP complex were determined by X-ray diffraction analysis to resolutions of 2.0 and 2.1 A, respectively. There is little observable change between these structures and that of wild-type RNase A, either free or with bound 3'-cytidine monophosphate. K7A/R10A/K66A RNase A was evaluated for its ability to cleave UpA, a dinucleotide substrate that does not span the P0 or the P2 subsites. In comparison to the wild-type enzyme, the value of kcat was decreased by 5-fold and that of kcat/Km was decreased 10-fold, suggesting that these remote subsites interact with the active site. These interactions were characterized by determining the pKa values of His12 and His119 at 0.018 and 0.142 M Na+, both in wild-type RNase A and the K7A/R10A/K66A variant. The side chains of Lys7, Arg10, and Lys66 depress the pKa values of these histidine residues, and this depression is sensitive to the salt concentration. In addition, the P0 and P2 subsites influence the interaction of His12 and His119 with each other, as demonstrated by changes in the cooperativity that gives rise to microscopic pKa values. Finally, the affinity of 3'-UMP for wild-type RNase A and the K7A/R10A/K66A variant at 0.018 and 0.142 M Na+ was determined by isothermal titration calorimetry. 3'-UMP binds to the variant protein with 5-fold weaker affinity at 0.018 M Na+ and 3-fold weaker affinity at 0.142 M Na+ than it binds to wild-type RNase A. Together these data demonstrate that long-range Coulombic interactions are an important feature in catalysis by RNase A.  相似文献   

7.
We report a study of the relative reactivity of the common amino acids and of their residues in lysozyme with osmium tetroxide, the osmium tetroxide-pyridine reagent, and with the oxo-osmium(VI)-pyridine reagent. With free amino acids, the osmium(VIII) reagents are most reactive with Met, Cys, His, Thr, Ser, Trp, Lys, and Pro; the osmium(VI) reagent only reacts significantly with His, Met, Cys, Thr, and Ser. In lysozyme, only Cys, Met, and Trp react extensively with the osmium(VIII) reagents; with the osmium(VI) reagent, Cys and Met are most reactive. We also note evidence both for cross-linking of proteins and for peptide bond cleavage, which appears to have considerable specificity for tryptophanyl residues.  相似文献   

8.
The leucine-to-alanine mutation at residue 201 of D-amino acid aminotransferase provides a unique enzyme which gradually loses its activity while catalyzing the normal transamination; the co-enzyme form is converted from pyridoxal 5'-phosphate to pyridoxamine 5'-phosphate upon the inactivation [Kishimoto,K., Yoshimura,T., Esaki,N., Sugio,S., Manning,J.M. and Soda,K. (1995) J. Biochem., 117, 691-696]. Crystal structures of both co-enzyme forms of the mutant enzyme have been determined at 2.0 A resolution: they are virtually identical, and are quite similar to that of the wild-type enzyme. Significant differences in both forms of the mutant are localized only on the bound co-enzyme, the side chains of Lys145 and Tyr31, and a water molecule sitting on the putative substrate binding site. Detailed comparisons of the structures of the mutant, together with that of the pyridoxamine-5'-phosphate form of the wild-type enzyme, imply that Leu201 would play a crucial role in the transamination reaction by keeping the pyridoxyl ring in the proper location without disturbing its oscillating motion, although the residue seems to not be especially important for the structural integrity of the enzyme.  相似文献   

9.
High-resolution crystallographic data show that Glu 168 and Glu 211 lie on opposite surfaces of the active site from Lys 345. Two different proposals for general base catalysis have emerged from these structural studies. In one scheme, the carboxylate side chains of Glu 168 and Glu 211 are proposed to ionize a trapped water molecule and the OH- serves as the base [Lebioda, L., & Stec, B. (1991) Biochemistry 30, 2817-2822]. In the other proposal, the epsilon-amino group of Lys 345 functions in general base catalysis [Wedekind, J. E., Poyner, R. R., Reed, G. H., & Rayment, I. (1994) Biochemistry 33, 9333-9342]. Genes encoding site specific mutations of these active site residues of yeast enolase, K345A, E168Q, and E211Q, have been prepared. The respective protein products of the wild type and mutant genes were expressed in Escherichia coli and isolated in homogeneous form. All three mutant proteins possess severely depressed activities in the overall reaction- < 1 part in 10(5) of wild type activity. Properties of the three mutant proteins in partial reactions were examined to define more clearly the roles of these residues in the catalytic cycle. The K345A variant fails to catalyze the exchange of the C-2 proton of 2-phospho-D-glycerate with deuterium in D2O, whereas both the E211Q and E168Q mutant proteins are functional in this partial reaction. For E211Q and E168Q enolases, exchange is essentially complete prior to appearance of product, and this observation provides further support for an intermediate in the normal reaction. K345A enolase is inactive in the ionization of tartronate semialdehyde phosphate (TSP), whereas both E168Q and E211Q proteins alter the tautomeric state or catalyze ionization of bound TSP. Wild type enolase catalyzes hydrolysis of (Z)-3-chloro-2-phosphoenolpyruvate by addition of OH- and elimination of Cl- at C-3. This reaction mimics the addition of OH- to C-3 of phosphoenolpyruvate in the reverse reaction with the normal product. All three mutant proteins are depressed in their abilities to carry out this reaction. In single-turnover assays, the activities vary in the order K345A > E168Q > E211Q. These results suggest that Lys 345 functions as the base in the ionization of 2-PGA and that Glu 211 participates in the second step of the reaction.  相似文献   

10.
Neuronal bungarotoxin (NBT) is a highly selective, slowly reversible, competitive antagonist of the alpha3beta2 neuronal nicotinic receptor. Contributions to NBT sensitivity are made by both the alpha3 and beta2 subunits. We used a chimeric alpha subunit to demonstrate that the entire alpha3 contribution lies within sequence segment 84-215. Construction and analysis of a series of mutant alpha3 subunits identified seven amino acid residues (Thr143, Tyr184, Lys185, His186, Ile188, Gln198, Ser203) within this region that contribute to NBT sensitivity. Changing Thr143 to lysine, as in alpha2, resulted in a approximately 1000-fold loss of NBT sensitivity. The effect on NBT sensitivity of changing each of the other six residues ranged from 1.8- to 40.5-fold. More extensive mutagenesis demonstrated that Thr143 serves as part of the consensus sequence for glycosylation at N141, and it is this glycosylation that is the determinant of NBT sensitivity. Only serine could substitute for threonine to maintain full NBT sensitivity, and changing Asn141 to alanine resulted in a approximately 300-fold loss of NBT sensitivity. The chimera alpha2-181-alpha3, containing all identified determinants except the glycosylation site, formed receptors insensitive to 300 nM NBT. Installation of threonine to complete the glycosylation consensus site in this chimera conferred NBT sensitivity only 10-fold less than that of wild-type alpha3beta2. These seven determinants of NBT sensitivity are located in close proximity to a series of conserved residues that are common features of all nicotinic receptor binding sites.  相似文献   

11.
Klebsiella aerogenes urease possesses a dinuclear metallocenter in which two nickel atoms are bridged by carbamylated Lys217. To assess whether carbamate-specific chemistry is required for urease activity, site-directed mutagenesis and chemical rescue strategies were combined in efforts to place a carboxylate group at the location of this metal ligand. Urease variants with Lys217 replaced by Glu, Cys, and Ala (K217E, K217C/C319A, and K217A proteins) were purified, shown to be activated by incubation with small organic acids plus Ni(II), and structurally characterized. K217C/C319A urease possessed a second change in which Cys319 was replaced by Ala in order to facilitate efforts to chemically modify Cys217; however, this covalent modification approach did not produce active urease. Chemical rescue of the K217E, K217C/C319A, and K217A variants required 2, 2, and 10 h, respectively, to reach maximal activity levels. The highest activity generated [224 micromol of urea degraded.min-1.(mg of protein)-1, for K217C/C319A urease incubated with 500 mM formic acid and 10 mM Ni at pH 6.5] corresponded to 56% of that measured for in vitro activation of the wild-type apoprotein. While the K217E apoprotein showed minimal structural perturbations, the K217C/C319A apoprotein showed a disordering of some active site residues, and the K217A apoprotein revealed a repositioning of His219 to allow the formation of a hydrogen bond with Thr169, thus replacing the hydrogen bond between the amino group of Lys217 and Thr169 in the native enzyme. Importantly, these structures allow rationalization of the relative rates and yields of chemical rescue experiments. The crystal structures of chemically rescued K217A and K217C/C319A ureases revealed a return of the active site residues to their wild-type positions. In both cases, noncovalently bound formate was structurally equivalent to the Lys-carbamate as the bridging metallocenter ligand. We conclude that carbamate-specific chemistry is not required for urease catalysis.  相似文献   

12.
The crystal structure of the mutant deoxyhemoglobin in which the beta-globin Val67(E11) has been replaced with threonine [Fronticelli et al. (1993) Biochemistry 32, 1235-1242] has been determined at 2.2 A resolution. Prior to the crystal structure determination, molecular modeling indicated that the Thr67(E11) side chain hydroxyl group in the distal beta-heme pocket forms a hydrogen bond with the backbone carbonyl of His63(E7) and is within hydrogen-bonding distance of the N delta of His63(E7). The mutant crystal structure indicates only small changes in conformation in the vicinity of the E11 mutation confirming the molecular modeling predictions. Comparison of the structures of the mutant beta-subunits and recombinant porcine myoglobin with the identical mutation [Cameron et al. (1993) Biochemistry 32, 13061-13070] indicates similar conformations of residues in the distal heme pocket, but there is no water molecule associated with either of the threonines of the beta-subunits. The introduction of threonine into the distal heme pocket, despite having only small perturbations in the local structure, has a marked affect on the interaction with ligands. In the oxy derivative there is a 2-fold decrease in O2 affinity [Fronticelli et al. (1993) Biochemistry 32, 1235-1242], and the rate of autoxidation is increased by 2 orders of magnitude. In the CO derivative the IR spectrum shows modifications with respect to that of normal human hemoglobin, suggesting the presence of multiple CO conformers. In the nitrosyl derivative an interaction with the O gamma atom of Thr67(E11) is probably responsible for the 10-fold increase in the rate of NO release from the beta-subunits. In the aquomet derivative there is a 6-fold decrease in the rate of hemin dissociation suggesting an interaction of the Fe-coordinated water with the O gamma of Thr67(E11).  相似文献   

13.
14.
Three trials were carried out with pigs between 5 and 8 wk of age to determine the limiting order of amino acids in a 13.5% CP corn-soybean meal-based diet containing 8% dried whey. The positive-control diet was a 19.2% CP corn-soybean meal-based diet (1.15% lysine), also with 8% dried whey. Amino acid additions to the low-protein, negative-control diet were based on levels needed to accomplish 110% of ideal ratios (to lysine, set at 1.15%). In Exp. 1, the addition of an amino acid mixture containing Lys, Trp, Thr, Met, Ile, and Val to the low-protein diet increased (P<.05) gain and gain: feed ratio, and these response traits were not different from those of pigs fed the 19.2% CP positive-control diet. Single deletion of Lys from the supplemental amino acid mixture depressed performance to a greater (P<.05) extent than single deletion of any of the other amino acids. Single deletions of Trp, Thr, Met, or Val decreased (P<.05) performance in a similar but lesser magnitude than the decrease caused by Lys deletion, whereas Ile deletion was without effect. Experiments 2 and 3 were designed to evaluate the limiting order of AA beyond Lys in the low-protein diet. Neither His nor Glu were found to be deficient, and, as in Exp. 1, deletion of Trp, Thr, Met, or Val from the supplemental amino acid mixture resulted in performance depressions (P<.05) that were similar. The results suggest that Lys is first-limiting and Trp, Thr, Met, and Val are equally second-limiting in a reduced protein (13.5% CP) corn-soybean meal-based diet with 8% whey for 10-kg pigs.  相似文献   

15.
Illumination of bovine pancreatic ribonuclease A (RNase A) in solution in the presence of rose bengal as a photosensitizer resulted in the progressive formation of enzyme dimers, trimers, tetramers and higher oligomers, as measured by gel electrophoresis and size exclusion chromatography. Oxygen was necessary for crosslink formation, and azide inhibition studies indicated that singlet oxygen was involved in the process. Chemical modification of His residues (with diethyl pyrocarbonate) and/or Lys residues (with acetic acid N-hydroxysuccinimide ester) in the enzyme decreased crosslinking, suggesting the participation of these two amino acid residues in the reaction. Met and cystine residues did not appear to be involved. Similar studies have shown that model N-(2-hydroxypropyl)methacrylamide (HPMA) copolymers containing epsilon-aminocaproic acid side chains terminating in His or Lys residues are photodynamically crosslinked via His-His or His-Lys interactions. Treatment of crosslinked RNase A and its His, Lys and Lys-His derivatives for 5 min at 97 degrees C in a dithiothreitol-sodium dodecyl sulfate mixture efficiently ruptured a major part of the photodynamically formed crosslinks; treatment with the detergent alone had no effect. Similar results were obtained with the crosslinked amino acid-containing HPMA copolymers, suggesting that photodynamic crosslinks involving His-His and His-Lys interaction are chemically the same in RNase A and the copolymer model.  相似文献   

16.
Structural studies of phospholipase C delta1 (PLCdelta1) in complexes with the inositol-lipid headgroup and calcium identified residues within the catalytic domain that could be involved in substrate recognition, calcium binding, and catalysis. In addition, the structure of the PLCdelta1 catalytic domain revealed a cluster of hydrophobic residues at the rim of the active site opening (hydrophobic ridge). To assess a role of each of these residues, we have expressed, purified, and characterized enzymes with the point mutations of putative active site residues (His311, Asn312, Glu341, Asp343, His356, Glu390, Lys438, Lys440, Ser522, Arg549, and Tyr551) and residues from the hydrophobic ridge (Leu320, Phe360, and Trp555). The replacements of most active site residues by alanine resulted in a great reduction (1,000-200,000-fold) of PLC activity analyzed in an inositol lipid/sodium cholate mixed micelle assay. Measurements of the enzyme activity toward phosphatidylinositol, phosphatidylinositol 4-monophosphate, and phosphatidylinositol 4, 5-bis-phosphate (PIP2) identified Ser522, Lys438, and Arg549 as important for preferential hydrolysis of polyphosphoinositides, whereas replacement of Lys440 selectively affected only hydrolysis of PIP2. When PLC activity was analyzed at different calcium concentrations, substitutions of Asn312, Glu390, Glu341, and Asp343 resulted in a shift toward higher calcium concentrations required for PIP2 hydrolysis, suggesting that all these residues contribute toward Ca2+ binding. Mutational analysis also confirmed the importance of His311 ( approximately 20,000-fold reduction) and His356 ( approximately 6,000-fold reduction) for the catalysis. Mutations within the hydrophobic ridge, which had little effect on PIP2 hydrolysis in the mixed-micelles, resulted in an enzyme that was less dependent on the surface pressure when analyzed in a monolayer. This systematic mutational analysis provides further insights into the structural basis for the substrate specificity, requirement for Ca2+ ion, catalysis, and surface pressure/activity dependence, with general implications for eukaryotic phosphoinositide-specific PLCs.  相似文献   

17.
The membrane-anchored glycoprotein CD59 inhibits assembly of the C5b-9 membrane attack complex (MAC) of human complement. This inhibitory function of CD59 is markedly selective for MAC assembled from human complement components C8 and C9, and CD59 shows little inhibitory function toward MAC assembled from rabbit and many other non-primate species. We have used this species selectivity of CD59 to identify the residues regulating its complement inhibitory function: cDNA of rabbit CD59 was cloned and used to express human/rabbit CD59 chimeras in murine SV-T2 cells. Plasma membrane expression of each CD59 chimera was quantified by use of a 5'-TAG peptide epitope, and each construct was tested for its ability to inhibit assembly of functional MAC from human versus rabbit C8 and C9. These experiments revealed that the species selectivity of CD59 is entirely determined by sequence contained between residues 42 and 58 of the human CD59 polypeptide, whereas chimeric substitution outside this peptide segment has little effect on the MAC inhibitory function of CD59. Substitution of human CD59 residues 42-58 into rabbit CD59 resulted in a molecule that was functionally indistinguishable from native human CD59, whereas the complementary construct (corresponding residues of rabbit CD59 substituted into human CD59) was functionally indistinguishable from rabbit CD59. Based on the solved solution structure of CD59, these data suggest that selectivity for human C8 and C9 resides in a cluster of closely spaced side chains on the surface of CD59 contributed by His44, Asn48, Asp49, Thr51, Thr52, Arg55, and Glu58 of the polypeptide.  相似文献   

18.
Tyrosine O-sulfation, a common post-translational modification in eukaryotes, is mediated by Golgi enzymes that catalyze the transfer of the sulfuryl group from 3'-phosphoadenosine 5'-phosphosulfate to tyrosine residues in polypeptides. We recently isolated cDNAs encoding human and mouse tyrosylprotein sulfotransferase-1 (Ouyang, Y. B., Lane, W. S., and Moore, K. L. (1998) Proc. Natl. Acad. Sci. U. S. A. 95, 2896-2901). Here we report the isolation of cDNAs encoding a second tyrosylprotein sulfotransferase (TPST), designated TPST-2. The human and mouse TPST-2 cDNAs predict type II transmembrane proteins of 377 and 376 amino acid residues, respectively. The cDNAs encode functional N-glycosylated enzymes when expressed in mammalian cells. In addition, preliminary analysis indicates that TPST-1 and TPST-2 have distinct specificities toward peptide substrates. The human TPST-2 gene is on chromosome 22q12.1, and the mouse gene is in the central region of chromosome 5. We have also identified a cDNA that encodes a TPST in the nematode Caenorhabditis elegans that maps to the right arm of chromosome III. Thus, we have identified two new members of a class of membrane-bound sulfotransferases that catalyze tyrosine O-sulfation. These enzymes may catalyze tyrosine O-sulfation of a variety of protein substrates involved in diverse physiologic functions.  相似文献   

19.
Since heparin potentiates activated factor XI (FXIa) inhibition by protease nexin-2 by providing a template to which both proteins bind (Zhang, Y., Scandura, J. M., Van Nostrand, W. E., and Walsh, P. N. (1997) J. Biol. Chem. 272, 26139-26144), we examined binding of factor XI (FXI) and FXIa to heparin. FXIa binds to heparin (Kd approximately 0.7 x 10(-9) M) >150-fold more tightly than FXI (Kd approximately 1.1 x 10(-7) M). To localize the heparin-binding site on FXI, rationally designed conformationally constrained synthetic peptides were used to compete with 125I-FXI binding to heparin. A peptide derived from the Apple 3 (A3) domain of FXI (Asn235-Arg266) inhibited FXI binding to heparin (Kd approximately 3.4 x 10(-6) M), whereas peptides from the A1 domain (Phe56-Ser86), A2 domain (Ala134-Ala176), and A4 domain (Ala317-Gly350) had no such effect. The recombinant A3 domain (rA3, Ala181-Val271) inhibited FXI binding to heparin (Ki approximately 1.4 x 10(-7) M) indicating that all the information necessary for FXI binding to heparin is contained entirely within the A3 domain. The A3 domain also contains a platelet-binding site (Asn235-Arg266), consisting of three surface-exposed loop structures, Pro229-Gln233, Thr741-Leu246, and Thr249-Phe260 (Baglia, F. A., Jameson, B. A., and Walsh, P. N. (1995) J. Biol. Chem. 270, 6734-6740). Only peptide Thr249-Phe260 (which contains a heparin binding consensus sequence, RIKKSKA) inhibits FXI binding to heparin (Ki = 2.1 x 10(-7) M), whereas peptides Pro229-Gln233 and Thr241- Leu246 had no effect. Fine mapping of the heparin-binding site using prekallikrein analogue amino acid substitutions of the synthetic peptide Thr249-Phe260 and alanine scanning of the recombinant A3 indicated that the amino acids Lys252 and Lys253 are important for heparin binding. Thus, the sequence Thr249-Phe260 which contains most of the binding energy for FXI interaction with platelets also mediates the binding of FXI to heparin.  相似文献   

20.
Peptidylglycine alpha-hydroxylating monooxygenase (PHM) is a copper, ascorbate, and molecular oxygen dependent enzyme that catalyzes the first step leading to the C-terminal amidation of glycine-extended peptides. The catalytic core of PHM (PHMcc), refined to residues 42-356 of the PHM protein, was expressed at high levels in CHO (DG44) (dhfr-) cells. PHMcc has 10 cysteine residues involved in 5 disulfide linkages. Endoprotease Lys-C digestion of purified PHMcc under nonreducing conditions cleaved the protein at Lys219, indicating that the protein consists of separable N- and C-terminal domains with internal disulfide linkages, that are connected by an exposed linker region. Disulfide-linked peptides generated by sequential CNBr and pepsin treatment of radiolabeled PHMcc were separated by reverse phase HPLC and identified by Edman degradation. Three disulfide linkages occur in the N-terminal domain (Cys47-Cys186, Cys81-Cys126, and Cys114-Cys131), along with three of the His residues critical to catalytic activity (His107, His108, and His172). Two disulfide linkages (Cys227-Cys334 and Cys293-Cys315) occur in the C-terminal domain, along with the remaining two essential His residues (His242, His244) and Met314, thought to be essential in binding one of the two nonequivalent copper atoms. Substitution of Tyr79 or Tyr318 with Phe increased the Km of PHM for its peptidylglycine substrate without affecting the Vmax. Replacement of Glu313 with Asp increased the Km 8-fold and decreased the kcat 7-fold, again identifying this region of the C-terminal domain as critical to catalytic activity. Taking into account information on the copper ligands in PHM, we propose a two-domain model with a copper site in each domain that allows spatial proximity between previously described copper ligands and residues identified as catalytically important.  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