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 Ras and Rap1A proteins can bind to the Raf and RalGDS families. Ras and Rap1A have Glu and Lys, respectively, at position 31. In the present study, we analyzed the effects of mutating the Glu at position 31 of the c-Ha-Ras protein to Asp, Ala, Arg, and Lys on the interactions with Raf-1 and RalGDS. The Ras-binding domain (RBD) of Raf-1 binds the E31R and E31K Ras mutants less tightly than the wild-type, E31A, and E31D Ras proteins; the introduction of the positively charged Lys or Arg residue at position 31 specifically impairs the binding of Ras with the Raf-1 RBD. On the other hand, the ability of the oncogenic RasG12V protein to activate Raf-1 in HEK293 cells was only partially reduced by the E31R mutation but was drastically impaired by the E31K mutation. Correspondingly, RasG12V(E31K) as well as Rap1A, but not RasG12V(E31R), exhibited abnormally tight binding with the cysteine-rich domain of Raf-1. On the other hand, the E31A, E31R, and E31K mutations, but not the E31D mutation, enhanced the RalGDS RBD-binding activity of Ras, indicating that the negative charge at position 31 of Ras is particularly unfavorable to the interaction with the RalGDS RBD. RasG12V(E31K), RasG12V(E31A), and Rap1A stimulate the RalGDS action more efficiently than the wild-type Ras in the liposome reconstitution assay. All of these results clearly show that the sharp contrast between the characteristics of Ras and Rap1A, with respect to the interactions with Raf-1 and RalGDS, depends on their residues at position 31.  相似文献   

2.
We have identified mutations in Raf-1 that increase binding to Ras. The mutations were identified making use of three mutant forms of Ras that have reduced Raf-1 binding (Winkler, D. G., Johnson, J. C., Cooper, J. A., and Vojtek, A. B. (1997) J. Biol. Chem. 272, 24402-24409). One mutation in Raf-1, N64L, suppresses the Ras mutant R41Q but not other Ras mutants, suggesting that this mutation structurally complements the Ras R41Q mutation. Missense substitutions of residues 143 and 144 in the Raf-1 cysteine-rich domain were isolated multiple times. These Raf-1 mutants, R143Q, R143W, and K144E, were general suppressors of three different Ras mutants and had increased interaction with non-mutant Ras. Each was slightly activated relative to wild-type Raf-1 in a transformation assay. In addition, two mutants, R143W and K144E, were active when tested for induction of germinal vesicle breakdown in Xenopus oocytes. Interestingly, all three cysteine-rich domain mutations reduced the ability of the Raf-1 N-terminal regulatory region to inhibit Xenopus oocyte germinal vesicle breakdown induced by the C-terminal catalytic region of Raf-1. We propose that a direct or indirect regulatory interaction between the N- and C-terminal regions of Raf-1 is reduced by the R143W, R143Q, and K144E mutations, thereby increasing access to the Ras-binding regions of Raf-1 and increasing Raf-1 activity.  相似文献   

3.
Antibody NC41 binds to the subtype N9 neuraminidase (NA) of influenza virus A/tern/Australia/ G70c/75 and inhibits its enzyme activity. To address the molecular mechanisms by which antibodies interact with neuraminidase and the requirements for successful escape from antibody inhibition, we made amino acid substitutions in heavy chain CDRs of NC41. Antibody proteins expressed as a single-chain Fv (scFv) fused with maltose-binding protein were assayed for binding to NA by ELISA. Association constants (Ka) for wild-type and mutant scFvs are as follows: wild type, 2 x 10(7) M-1; Asn31-->Gln, 2 x 10(7) M-1; Glu96-->Asp, 1 x 10(7) M-1; Asp97-->Lys, 6 x 10(6) M-1; and Asn98-->Gln, 8 x 10(6) M-1. The Ka for intact NC41 antibody was 4 x 10(8) M-1 in the same assay, reflecting increased stability compared to that of the scFv. Mutations in the scFv antibody had less of an effect on binding than mutations in their partners on the NA, and modeling studies suggest that interactions involving the mutant antibody side chains occur, even without taking increased flexibility into account. Asp97 forms a salt link with NA critical contact Lys434; of the four mutants, D97K shows the largest reduction in binding to NA. Mutant N98Q also shows reduced binding, most likely through the loss of interaction with NA residue Thr401. Substitution N31Q had no effect on Ka. NC41 residue Glu96 interacts with NA critical contact Ser368, yet E96D showed only a 2-fold reduction in binding to NA, apparently because the H bond can still form. Asp97 and Asn98 provide the most important interactions, but some binding is maintained when they are mutated, in contrast to their partners on the NA. The results are consistent with maturation of the immune response, when the protein epitope is fixed while variation in the antibody paratope allows increasing affinity. Influenza viruses may exploit this general mechanism since single amino acid changes in the epitope allow the virus to escape from the antibody.  相似文献   

4.
The serine protease trypsin was converted into a site-specific protease which hydrolyzes peptides between dibasic residues. Trypsin exhibits a high S1 specificity for Arg and Lys residues. However, the S1' specificity of trypsin is very broad, with only a slight preference for hydrophobic residues in P1'. We replaced Lys60 with Glu and Asp to introduce a high specificity for basic residues into the S1' site of trypsin. Both mutations cause a dramatic increase in the S1' specificity for Arg and Lys as measured by acyl transfer reactions. In K60E, the preference for Arg increases 70-fold while the preference for P1'-Lys increases 12-fold. In contrast, the preferences for other P1' residues either decrease slightly or remain the same. Thus, K60E prefers P1'-Arg over most other P1' residues by 2 orders of magnitude. Similar results are obtained when P1' specificity is measured in peptide cleavage assays. K60D exhibits an S1' specificity profile very similar to that of K60E, although the P1'-Arg preference is reduced by a factor of 2.5. Molecular modeling studies suggest that the high S1' specificity for Arg in K60E may be due to the formation of a salt bridge between Glu60 and the P1'-Arg of the substrate.  相似文献   

5.
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   

6.
We employed site-directed mutagenesis based on sequence comparisons and characterization of purified mutant enzymes to identify Glu558 and Asp766 of Syrian hamster 3-hydroxy-3-methylglutaryl coenzyme A (HMG-CoA) reductase (EC 1.1.1.34) as essential for catalysis. Mutant enzymes E558D, E558Q, and D766N had wild-type Km values for (S)-HMG-CoA and NADPH, but exhibited less than 0.5% of the wild-type catalytic activity. The inactive mutant polypeptides E558Q and D766N nevertheless can associate to generate an active enzyme. In vitro, 6% of the wild-type activity was observed when mutant polypeptides E558D and D766N were mixed in the absence of chaotropic agents. When mutant polypeptides E558Q and D766N were co-expressed in Escherichia coli, the resulting purified enzyme had 25% of wild-type activity. Hamster HMG-CoA reductase thus is a two-site, dimeric enzyme whose subunits associate to form an active site in which each monomer contributes at least one residue (e.g. Glu558 from one monomer and Asp766 from the other). The wild-type enzyme behaves as a dimer during size exclusion chromatography and has one HMG-CoA binding site per monomer. Syrian hamster HMG-CoA reductase thus appears to be a homodimer with two active sites which are located at the subunit interface.  相似文献   

7.
The serine/threonine kinase Raf-1 functions downstream from Ras to activate mitogen-activated protein kinase kinase, but the mechanisms of Raf-1 activation are incompletely understood. To dissect these mechanisms, wild-type and mutant Raf-1 proteins were studied in an in vitro system with purified plasma membranes from v-Ras- and v-Src-transformed cells (transformed membranes). Wild-type (His)6- and FLAG-Raf-1 were activated in a Ras- and ATP-dependent manner by transformed membranes; however, Raf-1 proteins that are kinase defective (K375M), that lack an in vivo site(s) of regulatory tyrosine (YY340/341FF) or constitutive serine (S621A) phosphorylation, that do not bind Ras (R89L), or that lack an intact zinc finger (CC165/168SS) were not. Raf-1 proteins lacking putative regulatory sites for an unidentified kinase (S259A) or protein kinase C (S499A) were activated but with apparently reduced efficiency. The kinase(s) responsible for activation by Ras or Src may reside in the plasma membrane, since GTP loading of plasma membranes from quiescent NIH 3T3 cells (parental membranes) induced de novo capacity to activate Raf-1. Wild-type Raf-1, possessing only basal activity, was not activated by parental membranes in the absence of GTP loading. In contrast, Raf-1 Y340D, possessing significant activity, was, surprisingly, stimulated by parental membranes in a Ras-independent manner. The results suggest that activation of Raf-1 by phosphorylation may be permissive for further modulation by another membrane factor, such as a lipid. A factor(s) extracted with methanol-chloroform from transformed membranes or membranes from Sf9 cells coexpressing Ras and SrcY527F significantly enhanced the activity of Raf-1 Y340D or active Raf-1 but not that of inactive Raf-1. Our findings suggest a model for activation of Raf-1, wherein (i) Raf-1 associates with Ras-GTP, (ii) Raf-1 is activated by tyrosine and/or serine phosphorylation, and (iii) Raf-1 activity is further increased by a membrane cofactor.  相似文献   

8.
Crystal structures have been determined of recombinant human tumor necrosis factor-alpha (TNF-alpha) and its R31D mutant that preferentially binds to TNF receptor R1 with more than seven times the relative affinity of binding to receptor R2. Crystals of the wild-type TNF were of space group P4(1)2(1)2 and had unit cell dimensions of a = b = 94.7 and c = 117.4 A. Refinement of the structure gave an R-factor of 22.3% at 2.5 A resolution. The crystals of TNF R31D mutant diffracted to 2.3 A resolution, and were of identical space group to the wild type with unit cell dimensions of a = b = 95.4 and c = 116.2 A, and the structure was refined to an R-factor of 21.8%. The trimer structures of the wild-type and mutant TNF were similar with a root mean square (r.m.s.) deviation of 0.56 A for Calpha atoms; however, the subunits within each trimer were more variable with an average r.m.s. deviation of 1.00 A on Calpha atoms for pairwise comparison of subunits. Model complexes of TNF with receptors R1 and R2 have been used to predict TNF-receptor interactions. Arg31 in all three subunits of wild-type TNF is predicted to form an ionic interaction with the equivalent glutamic acid in both receptors R1 and R2. Asp31 of the TNF R31D mutant is predicted to interact differently with the two receptors. The side chain of Asp31 in two subunits of the TNF mutant is predicted to form hydrogen bond interactions with Ser59 or Cys70 of R1, while it has no predicted interactions with R2. The loss of three strong ionic interactions of Arg31 and the electrostatic repulsion of Asp31 with Glu in the receptors is consistent with the reduced binding of the R31D mutant to both receptors relative to wild-type TNF. The replacement of these ionic interactions by two weaker hydrogen bond interactions between Asp31 of the R31D mutant and R1, compared with no interactions with R2, is in agreement with the observed preferential binding of the R31D mutant to R1 over R2. Analysis of the structure and function of receptor-discriminating mutants of TNF will help understand the biological role of TNF and facilitate its use as an antitumor agent.  相似文献   

9.
A carboxyl group of Asp-285 is essential for tetracycline/H+ antiport mediated by the transposon Tn10-encoded metal-tetracycline/H+ antiporter (TetA) of Escherichia coli (Yamaguchi, A., Akasaka, T., Ono, N., Someya, Y., Nakatani, M., and Sawai, T. (1992) J. Biol. Chem. 267, 7490-7498). Spontaneous tetracycline resistance revertants were isolated from E. coli cells carrying the Asn-285 mutant tetA gene. All of the revertants were due to the second-site mutation at codon 220 of GCG (Ala) to GAG (Glu). The Km value of the tetracycline transport mediated by the revertant TetA protein was about 4-fold higher than that of the wild-type, indicating that the revertant is a low affinity mutant. A Glu-220 and Asn-285 double mutant constructed by site-directed mutagenesis showed the same properties as the revertants, confirming that the mutation of Ala-220 is solely responsible for the suppression. The Asp-220 mutation of the Asn-285 mutant resulted in a lower level of restoration of the tetracycline resistance and the transport activity than in the case of the Glu-220 mutation. A single mutation replacing Ala-220 with Glu or Asp caused about a 2-4-fold decrease in the tetracycline resistance, but no crucial change in the transport activity. It is not likely that Glu-220 is required for a charge-neutralizing salt bridge because an unpaired negative charge in a Glu-220 or Asp-220 single mutant did not cause a serious change in the activity. An alternative explanation is reasonable; Asp-285 directly contributes to the binding of a cationic substrate, metal-tetracycline chelation complex, or proton, and an acidic residue at position 220 can take over the role of Asp-285.  相似文献   

10.
The crystal structure of rabbit muscle pyruvate kinase complexed with Mn2+, K+, and pyruvate revealed a binding site of K+ [T. M. Larsen, L. T. Laughlin, H. M. Holden, I. Rayment, and G. H. Reed (1994) Biochemistry 33, 6301-6309]. Sequence comparisons of rabbit muscle pyruvate kinase and pyruvate kinases from Corynebacterium glutamicum and Escherichia coli, which do not exhibit a requirement for activation by monovalent cations, indicate that the only substitutions in the K+ binding site are conservative. Glu 117 in the rabbit muscle enzyme, which is close to the K+ site, is, however, replaced by Lys in these two bacterial pyruvate kinases. The proximity of Glu 117 to K+ in the structure of the rabbit enzyme and conservation of the binding site in the bacterial enzymes which lack a dependence on monovalent cations suggested that a protonated epsilon-amino group of Lys 117 in these bacterial enzymes may provide an "internal monovalent cation." Site-specific mutant forms of the rabbit enzyme corresponding to E117K, E117A, E117D, and E117K/K114Q pyruvate kinase were examined to test this hypothesis. The E117K pyruvate kinase exhibits 12% of the activity of the fully activated wild-type enzyme but is > 200-fold more active than the wild-type enzyme in the absence of activating monovalent cations. Moreover, the activity of E117K pyruvate kinase exhibits no stimulation by monovalent cations in the assay mixtures. Both E117A and E117D pyruvate kinases retain activation by monovalent cations but have reduced activities relative to wild type. The results are consistent with the hypothesis that pyruvate kinases that do not require activation by monovalent cations supply an internal monovalent cation in the form of a protonated epsilon-amino group of Lys. The results also support the assignment of the monovalent cation in the active site of pyruvate kinase.  相似文献   

11.
Activation of cytoplasmic serine/threonine kinase Raf-1, an important effector of Ras, requires direct binding to Ras. The yeast two-hybrid screening system used for identification of inhibitors of Ras/Raf-1 interaction showed radicicol to be an inhibitor. Radicicol has been shown to induce morphological reversion of transformed cells. Immunoprecipitation with an anti-Ras antibody revealed that the in vivo Ras/Raf-1 binding in v-Ha-ras-transformed cells was also blocked by low concentrations of radicicol (0.1 approximately 1 microg/ml), while degradation of Raf-1 was induced at concentrations higher than 2 microg/ml. However, in vitro binding of glutathion S-transferase-fused Ras to a maltose binding protein-fused RIP3 containing the Ras-binding domain (RBD) of Raf-1 was not inhibited by radicicol. Similar two-hybrid assays with several truncated forms of Raf-1 showed that both the conserved serine/threonine-rich domain (CR2) and the C-terminal protein kinase domain (CR3) were required for the full inhibition by radicicol. These results suggest that radicicol interacts directly or indirectly with the region except with RBD of Raf-1, thereby inhibiting a conformational change of Raf-1 prerequisite for binding to Ras.  相似文献   

12.
Our goal was to identify pairs of charged residues in the membrane domains of the Na+/glucose cotransporter (SGLT1) that form salt bridges, to obtain information about packing of the transmembrane helices. The strategy was to neutralize Glu225, Asp273, Asp294, and Lys321 in helices 6-8, express the mutants in oocytes, measure [14C]-alphaMDG uptake, and then attempt to find second-site mutations of opposite charge that restored function. alphaMDG uptake by E225A was identical to that by SGLT1, whereas transport was reduced by over 90% for D273A, D294A, and K321A and was not restored in the double mutants D273A/K321A or D294A/K321A. This suggested that K321 did not form salt bridges with D273 or D294 and that E225 was not involved in salt-bridging. Neutralization of K321 dramatically changed the Na+ uniport and Na+/glucose cotransport kinetics. The maximum rate of uniport in K321A increased 3-5-fold with a decrease in the apparent affinity for Na+ (70 vs 3 mM) and no change in apparent H+ affinity (0.5 microM). The change in Na+ affinity caused a +50 mV shift in the charge/voltage (Q/V) and relaxation time constant (tau)/voltage curves in the presteady-state kinetics. The presteady-state kinetics in H+ remained unchanged. The lower Na+ affinity resulted also in a 200-fold increase in the apparent K0.5 for alphaMDG and phlorizin. Replacements of K321 with alanine, valine, glutamine, arginine, or glutamic acid residues changed the steady-state kinetics in a similar way. Therefore, we suggest that K321 determines, directly or indirectly, (i) the rate and selectivity of SGLT1 uniport activity and (ii) the apparent affinities of SGLT1 for Na+, and indirectly sugar in the cotransport mode.  相似文献   

13.
Cytotoxic T lymphocyte (CTL) activation requires specific T cell receptor (TCR)-class I major histocompatibility complex (MHC) antigen complex interactions as well as the participation of coreceptor or accessory molecules on the surface of CTL. CD8 can serve as a coreceptor in that it binds to the same MHC class I molecules as the TCR to facilitate efficient TCR signaling. In addition, CD8 can be "activated" by TCR stimulation to bind to class I molecules with high avidity, including class I not recognized by the TCR as antigenic complexes (non-antigen [Ag] class I), to augment CTL responses and thus serve an accessory molecule function. A Glu/Asp227-->Lys substitution in the class I alpha 3 domain acidic loop abrogates lysis of target cells expressing these mutant molecules by alloreactive CD8-dependent CTL. Lack of response is attributed to the destruction of the CD8 binding site in the alpha 3 domain which is likely to disrupt CD8 coreceptor function. The relative importance of the class I alpha 3 domain acidic loop Glu227 in coreceptor as opposed to accessory functions of CD8 is unclear. To address this issue, we examined CTL adhesion and degranulation in response to immobilized class I-peptide complexes formed in vitro from antigenic peptides and purified class I molecules containing wild-type or Glu227-->Lys substituted alpha 3 domains. The alpha 3 domain mutant class I-peptide complexes were bound by CTL and triggered degranulation, however to much lower levels than wild-type class I-peptide complexes. In further experiments, it is directly demonstrated that the alpha 3 domain mutant class I molecules, which lack the Glu227 CD8 binding site, still serve as TCR-activated, avidity-enhanced CD8 accessory ligands. However, mutant class I peptide Ag complexes failed to effectively serve as CD8 coreceptor ligands to initiate TCR-dependent signals required to induce avidity-enhanced CD8 binding to coimmobilized non-Ag class I molecules. Thus the Glu227-->Lys mutation effectively distinguishes CD8 coreceptor and avidity-enhanced CD8 accessory functions.  相似文献   

14.
Pyridine nucleotide transhydrogenase (EC 1.6.1.1) from Escherichia coli was investigated with respect to the role of glutamic and aspartic acid residues reactive to N,N'-dicyclohexylcarbodiimide (DCCD) and potentially involved in the proton-pumping mechanism of the enzyme. The E. coli transhydrogenase consists of an alpha (510 residues) and a beta (462 residues) subunit. DCCD reacts with the enzyme to inhibit catalytic activity and proton pumping. This reagent modifies Asp alpha 232, Glu alpha 238, and Glu alpha 240 as well as amino acid residue(s) in the beta subunit. Using the cloned and overexpressed E. coli transhydrogenase genes (Clarke, D. M., and Bragg, P. D. (1985) J. Bacteriol. 162, 367-373), Asp alpha 232 and Glu alpha 238 were replaced independently by site-specific mutagenesis. In addition, Asp alpha 232, Glu alpha 238, and Glu alpha 240 were replaced to generate triple mutants. The specific catalytic activities of the mutant transhydrogenases alpha D232N, alpha D232E, alpha D232K, alpha D232H, alpha E238K, and alpha E238Q as well as of the triple mutants alpha D232N, alpha E238Q, alpha E240Q and alpha D232H, alpha E238Q, alpha E240Q were in the range of 40-90% of the wild-type activity. Proton-pumping activity was present in all mutants. Examination of the extent of subunit modification by [14C]DCCD revealed that the label was still incorporated into both alpha and beta subunits in the Asp alpha 232 mutants, but that the alpha subunit was not labeled in the triple mutants. Catalytic and proton-pumping activities were nearly insensitive to DCCD in the triple mutants. This suggests that loss of catalytic and proton-pumping activities is associated with modification of the aspartic and glutamic acid residues of the alpha subunit. In the presence of the substrate NADPH, the rate of modification of the beta subunit by [14C]DCCD was increased, and there was a greater extent of enzyme inactivation. By contrast, NADH and 3-acetylpyridine-NAD+ protected the catalytic activity of the transhydrogenase from inhibition by DCCD. The protection was particularly marked in the E238Q and E238K mutants. It is concluded that the Asp alpha 232, Glu alpha 238, and Glu alpha 240 residues are not essential for catalytic activity or proton pumping. The inactivation by DCCD is likely due to the introduction of a sterically hindering group that reacts with the identified acidic residues close to the NAD(H)-binding site.  相似文献   

15.
Mutagenesis of Glu820, present in the catalytic subunit of gastric H+,K+-ATPase, into an Asp hardly affects K+-stimulated ATPase and K+-stimulated dephosphorylation of the enzyme. The ATP phosphorylation rate of the E820D mutant, however, is rather low and the apparent affinity for ATP in the phosphorylation process of this mutant is 2-3 times lower than that of the wild type enzyme. The reduction in the ATP phosphorylation rate of the E820D mutant has only an effect on the ATPase activity at low temperature. These findings suggest that Glu820 might play a role in H+ stimulation of the phosphorylation process.  相似文献   

16.
We have expressed and characterized a mutant of Xenopus laevis Cu,Zn superoxide dismutase in which four highly conserved charged residues belonging to the electrostatic loop have been replaced by neutral side chains: Lys120 --> Leu, Asp130 --> Gln, Glu131 --> Gln, and Lys134 --> Thr. At low ionic strength, the mutant enzyme is one of the fastest superoxide dismutases ever assayed (k = 6.7 x 10(9) M(-1) s(-1), at pH 7 and mu = 0.02 M). Brownian dynamics simulations give rise to identical enzyme-substrate association rates for both wild-type and mutant enzymes, ruling out the possibility that enhancement of the activity is due to pure electrostatic factors. Comparative analysis of the experimental catalytic rate of the quadruple and single mutants reveals the nonadditivity of the mutation effects, indicating that the hyperefficiency of the mutant is due to a decrease of the energy barrier and/or to an alternative pathway for the diffusion of superoxide within the active site channel. At physiological ionic strength the catalytic rate of the mutant at neutral pH is similar to that of the wild-type enzyme as it is to the catalytic rate pH dependence. Moreover, mutation effects are additive. These results show that, at physiological salt conditions, electrostatic loop charged residues do not influence the diffusion pathway of the substrate and, if concomitantly neutralized, are not essential for high catalytic efficiency of the enzyme, pointing out the role of the metal cluster and of the invariant Arg141 in determining the local electrostatic forces facilitating the diffusion of the substrate towards the active site.  相似文献   

17.
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   

18.
The zymogen and mature enzyme forms of trypsin-like serine proteases exhibit a wide range of activities. The prototypical trypsinogen-trypsin system is an example of a minimally active zymogen and a maximally active mature protease. The present work identifies several features of trypsinogen which govern its activity. Our results indicate that rat trypsin is 10(8)-fold more active than rat trypsinogen. Rat trypsinogen appears to be less active than bovine trypsinogen. His40 is believed to be an important determinant of zymogen activity. We are unable to verify this role for His40 in trypsinogen since the mutation of His40 to Phe appears to change the trypsin-substrate interface. Deletion of the N-terminal Ile16 from trypsin is expected to produce a trypsinogen-like protein since the Ile16-Asp194 salt bridge cannot form. Such mutants have higher activity and BPTI affinity than trypsinogen, which indicates that the activation peptide stabilizes the inactive trypsinogen conformation. The mutation of Lys15 to Ala increases the BPTI affinity and activity of trypsinogen to an even greater extent; thus, removal of Lys15 can account for the effect of the loss of the activation peptide. These results suggest that Lys15 is an important determinant of zymogen activity. The mutation of Asp194 to Asn also increases the BPTI affinity and activity of trypsinogen. This result suggests that in addition to stabilizing the active conformation of trypsin via the Ile16-Asp194 salt bridge, Asp194 also maintains the inactive conformation of trypsinogen. A correlation exists between the values of kcat/Km and BPTI affinity of mutant trypsinogens and trypsins. However, the slope of this correlation is 0.64, which indicates that different "active" conformations are involved in BPTI binding and substrate hydrolysis. DeltaI16V17 trypsinogen is the lone outlier; its BPTI affinity is higher than would be expected based on the value of kcat/Km. We show that the rate of BPTI association is slower for DeltaI16V17 trypsinogen than for a mutant trypsinogen with a similar BPTI affinity. This observation suggests that BPTI binds to an "active" trypsinogen conformation that is not kinetically accessible to substrates.  相似文献   

19.
Site-directed mutagenesis and assay of Rb+ and Tl+ occlusion in recombinant Na,K-ATPase from yeast were combined to establish structure-function relationships of amino acid side chains involved in high-affinity occlusion of K+ in the E2[2K] form. The wild-type yeast enzyme was capable of occluding 2 Rb+ or Tl+ ions/ouabain binding site or alpha 1 beta 1 unit with high apparent affinity (Kd(Tl+) = 7 +/- 2 microM), like the purified Na,K-ATPase from pig kidney. Mutations of Glu327(Gln,Asp), Asp804(Asn, Glu), Asp808(Asn, Glu) and Glu779(Asp) abolished high-affinity occlusion of Rb+ or Tl+ ions. The substitution of Glu779 for Gln reduced the occlusion capacity to 1 Tl+ ion/alpha 1 beta 1-unit with a 3-fold decrease of the apparent affinity for the ion (Kd(Tl+) = 24 +/- 8 microM). These effects on occlusion were closely correlated to effects of the mutations on K0.5(K+) for K+ displacement of ATP binding. Each of the four carboxylate residues Glu327, Glu779, and Asp804 or Asp808 in transmembrane segments 4, 5, and 6 is therefore essential for high-affinity occlusion of K+ in the E2[2K] form. These residues either may engage directly in cation coordination or they may be important for formation or stability of the occlusion cavity.  相似文献   

20.
Glu126 and Arg144 in the lactose permease are indispensable for substrate binding and probably form a charge-pair [Venkatesan, P., and Kaback, H. R. (1998) Proc. Natl. Acad. Sci. U.S.A. 95, 9802-9807]. Mutants with Glu126-->Ala or Arg144-->Ala do not bind ligand or catalyze lactose accumulation, efflux, exchange, downhill lactose translocation, or lactose-induced H+ influx. In contrast, mutants with conservative mutations (Glu126-->Asp or Arg144-->Lys) exhibit drastically different phenotypes. Arg144-->Lys permease accumulates lactose slowly to low levels, but does not bind ligand or catalyze equilibrium exchange, efflux, or lactose-induced H+ influx. In contrast, Glu126-->Asp permease catalyzes lactose accumulation and lactose-induced H+ influx to wild-type levels, but at significantly lower rates. Surprisingly, however, no significant exchange or efflux activity is observed. Glu126-->Asp permease exhibits about a 6-fold increase in the Km for active transport relative to wild-type permease with a comparable Vmax. Direct binding assays using flow dialysis demonstrate that mutant Glu126-->Asp binds p-nitrophenyl-alpha,D-galactopyranoside. Indirect binding assays utilizing substrate protection against [14C]-N-ethylmaleimide labeling of single-Cys148 permease reveal an apparent Kd of 3-5 mM for lactose and 15-20 microM for beta, D-galactopyranosyl-1-thio-beta,D-galactopyranoside (TDG). The affinity of Glu126-->Asp/Cys148 permease for lactose is markedly decreased (Kd > 80 mM), while TDG affinity is altered to a much lesser extent (Kd ca. 80 microM). The results extend the conclusion that a carboxylate at position 126 and a guanidinium group at position 144 are irreplaceable for substrate binding and support the idea that Arg144 plays a major role in substrate specificity.  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