首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
Pharaonis halorhodopsin (phR) is an inward light-driven chloride ion pump from Natronobacterium pharaonis. In order to clarify the role of Ser-130(phR) residue which corresponds to Ser-115(shR) for salinarum hR on the anion-binding affinity, the wild-type and Ser-130 mutants substituted with Thr, Cys and Ala were expressed in E. coli cells and solubilized with 0.1% n-dodecyl beta-D-maltopyranoside The absorption maximum (lambda(max)) of the S130T mutant indicated a blue shift from that of the wild type in the absence and presence of chloride. For S130A, a large red shift (12 nm) in the absence of chloride was observed. The wild-type and all mutants showed the blue-shift of lambda(max) upon Cl(-) addition, from which the dissociation constants of Cl(-) were determined. The dissociation constants were 5, 89, 153 and 159 mM for the wild-type, S130A, S130T and S130C, respectively, at pH 7.0 and 25 degrees C. Circular dichroic spectra of the wild-type and the Ser-130 mutants exhibited an oligomerization. The present study revealed that the Ser-130 of N. pharaonis halorhodopsin is important for the chloride binding.  相似文献   

2.
Phospholamban (PLB) can be phosphorylated at Ser(16) by cyclic AMP-dependent protein kinase and at Thr(17) by Ca(2+)-calmodulin-dependent protein kinase during beta-agonist stimulation. A previous study indicated that mutation of S16A in PLB resulted in lack of Thr(17) phosphorylation and attenuation of the beta-agonist stimulatory effects in perfused mouse hearts. To further delineate the functional interplay between dual-site PLB phosphorylation, we generated transgenic mice expressing the T17A mutant PLB in the cardiac compartment of the null background. Lines expressing similar levels of T17A mutant, S16A mutant, or wild-type PLB in the null background were characterized in parallel. Cardiac myocyte basal mechanics and Ca(2+) kinetics were similar among the three groups. Isoproterenol stimulation was associated with phosphorylation of both Ser(16) and Thr(17) in wild-type PLB and Ser(16) phosphorylation in T17A mutant PLB, whereas there was no detectable phosphorylation of S16A mutant PLB. Phosphorylation of Ser(16) alone in T17A mutant PLB resulted in responses of the mechanical and Ca(2+) kinetic parameters to isoproterenol similar to those in wild-type myocytes, which exhibited dual-site PLB phosphorylation. However, those parameters were significantly attenuated in the S16A mutant myocytes. Thus, Ser(16) in PLB can be phosphorylated independently of Thr(17) in vivo, and phosphorylation of Ser(16) is sufficient for mediating the maximal cardiac responses to beta-adrenergic stimulation.  相似文献   

3.
Chloroplast glyceraldehyde-3-phosphate dehydrogenase (GAPDH) of higher plants uses both NADP(H) and NAD(H) as coenzyme and consists of one (GapA) or two types of subunits (GapA, GapB). AB-GAPDH is regulated in vivo through the action of thioredoxin and metabolites, showing higher kinetic preference for NADPH in the light than in darkness due to a specific effect on kcat(NADPH). Previous crystallographic studies on spinach chloroplast A4-GAPDH complexed with NADP or NAD showed that residues Thr33 and Ser188 are involved in NADP over NAD selectivity by interacting with the 2'-phosphate group of NADP. This suggested a possible involvement of these residues in the regulatory mechanism. Mutants of recombinant spinach GapA (A4-GAPDH) with Thr33 or Ser188 replaced by Ala (T33A, S188A and double mutant T33A/S188A) were produced, expressed in Escherichia coli, and compared to wild-type recombinant A4-GAPDH, in terms of crystal structures and kinetic properties. Affinity for NADPH was decreased significantly in all mutants, and kcat(NADPH) was lowered in mutants carrying the substitution of Ser188. NADH-dependent activity was unaffected. The decrease of kcat/Km of the NADPH-dependent reaction in Ser188 mutants resembles the behaviour of AB-GAPDH inhibited by oxidized thioredoxin, as confirmed by steady-state kinetic analysis of native enzyme. A significant expansion of size of the A4-tetramer was observed in the S188A mutant compared to wild-type A4. We conclude that in the absence of interactions between Ser188 and the 2'-phosphate group of NADP, the enzyme structure relaxes to a less compact conformation, which negatively affects the complex catalytic cycle of GADPH. A model based on this concept might be developed to explain the in vivo light-regulation of the GAPDH.  相似文献   

4.
The Gin residue at amino acid position 102 ofBacillus stearothermophilus lactate dehydrogenase was replaced with Ser, Thr, Tyr, or Phe to investigate the effect on substrate recognition. The Q102S and Q102T mutant enzymes were found to have a broader range of substrate specificity (measured byk cat/K m) than the wild-type enzyme. However, it is evident that either Ser or Thr at position 102 are of a size able to accommodate a wide variety of substrates in the active site and substrate specificity appears to rely largely on size discrimination in these mutants. The Q102F and Q102Y mutant enzymes have low catalytic efficiency and do not show this relaxed substrate specificity. However, their activities are restored by the presence of an aromatic substrate. All of the enzymes have a very low catalytic efficiency with branched chain aliphatic substrates.Abbreviations used BSLDH Bacillus stearothermophilus lactate dehydrogenase - FBP fructose-1,6-bisphosphate - HP hydroxypyruvate - KB ketobutyrate - KC ketocaproate - KV ketovalerate - MDH malate dehydrogenase - PP phenylpyruvate - PYR pyruvate - RBE relative binding energy  相似文献   

5.
Ser130, Asp131 and Asn132 ('SDN') are highly conserved residues in class A beta-lactamases forming one wall of the active-site cavity. All three residues of the SDN loop in Streptomyces albus G beta-lactamase were modified by site-directed mutagenesis. The mutant proteins were expressed in Streptomyces lividans, purified from culture supernatants and their kinetic parameters were determined for several substrates. Ser130 was substituted by Asn, Ala and Gly. The first modification yielded an almost totally inactive protein, whereas the smaller-side-chain mutants (A and G) retained some activity, but were less stable than the wild-type enzyme. Ser130 might thus be involved in maintaining the structure of the active-site cavity. Mutations of Asp131 into Glu and Gly proved to be highly detrimental to enzyme stability, reflecting significant structural perturbations. Mutation of Asn132 into Ala resulted in a dramatically decreased enzymic activity (more than 100-fold) especially toward cephalosporin substrates, kcat. being the most affected parameter, which would indicate a role of Asn132 in transition-state stabilization rather than in ground-state binding. Comparison of the N132A and the previously described N132S mutant enzymes underline the importance of an H-bond-forming residue at position 132 for the catalytic process.  相似文献   

6.
Sato M  Kanamori T  Kamo N  Demura M  Nitta K 《Biochemistry》2002,41(7):2452-2458
Pharaonis halorhodopsin (phR), the light-driven chloride ion pump from Natronobacterium pharaonis with C-terminal histidine tag, was expressed in Escherichia coli cells. The protein was solubilized with 0.1% n-dodecyl beta-D-maltopyranoside and purified with a nickel column. Removal of Cl- from the medium yields blue phR (phR(blue)) that has lost Cl- near the chromophore. Addition of Cl- converts phR(blue) to a red-shifted Cl--bound form (phR(Cl)). Circular dichroic spectra of phR(blue) and phR(Cl) exhibited a bilobe in the visual region, indicating specific oligomerization of the phR monomers. The order of anion concentration which induced a shift from phR(blue) to phR(X) was Br- < Cl- < NO3- < N3-, which was the same as in the case of phR purified from N. pharaonis membranes. Chloride binding kinetics was measured by time-resolved absorption changes with stopped-flow rapid mixing. Rates of Cl- binding consisted of fast and slow components, and the amplitude of the fast component was about 90% of the total changes. The rate constant of the fast component at 100 mM NaCl at 25 degrees C was 260 s(-1) with an apparent activation energy of 35 kJ/mol. These values are in good agreement with the process of Cl- uptake in the photocycle (O --> hR' reaction) reported previously [Váró et al. (1995) Biochemistry 34, 14500-14507]. In addition, the Cl- concentration dependence on both rates was similar to each other. These observations suggest that the O-intermediate is similar to phR(blue) and that Cl- uptake during the photocycle may be ruled by a passive process.  相似文献   

7.
The structures of three mutants of bacteriophage T4 lysozyme selected using a screen designed to identify thermostable variants are described. Each of the mutants has a substitution involving threonine. Two of the variants, Thr 26-->Ser (T26S) and Thr 151-->Ser (T151S), have increased reversible melting temperatures with respect to the wild-type protein. The third, Ala 93-->Thr (A93T), has essentially the same stability as wild type. Thr 26 is in the wall of the active-site cleft. Its replacement with serine results in the rearrangement of nearby residues, most notably Tyr 18, suggesting that the increase in stability may result from the removal of strain. Thr 151 in the wild-type structure is far from the active site and appears to sterically prevent the access of solvent to a preformed binding site. In the mutant, the removal of the methyl group allows access to the solvent binding site and, in addition, the Ser 151 hydroxyl rotates to a new position so that it also contributes to solvent binding. Residue 93 is in a highly exposed site on the surface of the molecule, and presumably is equally solvent exposed in the unfolded protein. It is, therefore, not surprising that the substitution Ala 93-->Thr does not change stability. The mutant structures show how chemically similar mutations can have different effects on both the structure and stability of the protein, depending on the structural context. The results also illustrate the power of random mutagenesis in obtaining variants with a desired phenotype.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

8.
To investigate the roles of PTEN (phosphatase and tensin homolog deleted on chromosome 10) in the regulation of 3-position phosphorylated phosphoinositide metabolism as well as insulin-induced Akt phosphorylation and glucose metabolism, wild-type PTEN and its phosphatase-dead mutant (C124S) with or without an N-terminal myristoylation tag were overexpressed in Sf-9 cells and 3T3-L1 adipocytes using baculovirus and adenovirus systems, respectively. When expressed in Sf-9 cells together with the p110alpha catalytic subunit of phosphoinositide 3-kinase, myristoylated PTEN markedly reduced the accumulations of both phosphatidylinositol 3,4-bisphosphate and phosphatidylinositol 3,4,5-trisphosphate induced by p110alpha. In contrast, overexpression of the C124S mutants apparently increased these accumulations. In 3T3-L1 adipocytes, insulin-induced accumulations of phosphatidylinositol 3,4-bisphosphate and phosphatidylinositol 3,4,5-trisphosphate were markedly suppressed by overexpression of wild-type PTEN with the N-terminal myristoylation tag, but not by that without the tag. On the contrary, the C124S mutants of PTEN enhanced insulin-induced accumulations of phosphatidylinositol 3,4-bisphosphate and phosphatidylinositol 3,4,5-trisphosphate. Interestingly, the phosphorylation level of Akt at Thr308 (Akt2 at Thr309), but not at Ser473 (Akt2 at Ser474), was revealed to correlate well with the accumulation of phosphatidylinositol 3,4,5-trisphosphate modified by overexpression of these PTEN proteins. Finally, insulin-induced increases in glucose transport activity were significantly inhibited by the overexpression of myristoylated wild-type PTEN, but were not enhanced by expression of the C124S mutant of PTEN. Therefore, in conclusion, 1) PTEN dephosphorylates both phosphatidylinositol 3,4-bisphosphate and phosphatidylinositol 3,4,5-trisphosphate in vivo, and the C124S mutants interrupt endogenous PTEN activity in a dominant-negative manner. 2) The membrane targeting process of PTEN may be important for exerting its function. 3) Phosphorylations of Thr309 and Ser474 of Akt2 are regulated differently, and the former is regulated very sensitively by the function of PTEN. 4) The phosphorylation level of Ser474, but not that of Thr309, in Akt2 correlates well with insulin-stimulated glucose transport activity in 3T3-L1 adipocytes. 5) The activity of endogenous PTEN may not play a major role in the regulation of glucose transport activity in 3T3-L1 adipocytes.  相似文献   

9.
Dong S  Ybe JA  Hecht MH  Spiro TG 《Biochemistry》1999,38(11):3379-3385
Type I Cu proteins maintain a trigonal N2S coordination group (with weak axial ligation) in both oxidation states of the Cu2+/+ ion, thereby reducing the reorganization energy for electron transfer. Requirements for maintaining this coordination group were investigated in poplar plastocyanin (Pcy) by mutation of a conserved element of the type 1 architecture, an asparagine residue (Asn38) adjacent to one of the ligating histidines. The side chain of this asparagine forms an active site clasp via two H-bonds with the residue (Ser85) adjacent to the ligating cysteine (Cys84). In addition, the main chain NH of Asn38 donates an H-bond to the thiolate ligand. We have investigated the importance of these interactions by mutating Asn38 to Gln, Thr, and Leu. The mutant proteins are capable of folding and binding Cu2+, but the blue color fades; the rate of fading increases in the order Gln < Thr < Leu. The color is not restored by ferricyanide, showing that the protein is modified irreversibly, probably by oxidation of Cys84. The more stable mutants N38Q and N38T were characterized spectroscopically. The wild-type properties are slightly perturbed for N38Q, but N38T shows remarkable similarity to another type 1 Cu protein, azurin (Azu) from Pseudomonas aeruginosa. The Cu-S(Cys) bond is longer in Azu than in Pcy, and the NH H-bond to the ligating S atom is shorter. Molecular modeling suggests a similar effect for N38T because the threonine residue shifts toward Ser85 in order to avoid a steric clash and to optimize H-bonding. These results demonstrate that H-bonding adjacent to the type 1 site stabilizes an architecture which both modulates the electronic properties of the Cu, and suppresses side reactions of the cysteine ligand.  相似文献   

10.
Sato M  Kubo M  Aizawa T  Kamo N  Kikukawa T  Nitta K  Demura M 《Biochemistry》2005,44(12):4775-4784
Natronomonas (Natronobacterium) pharaonis halorhodopsin (NpHR) is an inward light-driven Cl(-) ion pump. For efficient Cl(-) transport, the existence of Cl(-)-binding or -interacting sites in both extracellular (EC) and cytoplasmic (CP) channels is postulated. Candidates include Arg123 and Thr126 in EC channels and Lys215 and Thr218 in CP channels. The roles played by these amino acid residues in anion binding and in the photocycle have been investigated by mutation of the amino acid residues at these positions. Anion binding was assayed by changes in circular dichroism and the shift in the absorption maximum upon addition of Cl(-) to anion-free NpHR. The binding affinity was affected in mutants in which certain EC residues had been replaced; this finding revealed the importance of Arg123. On the other hand, mutants in which certain residues in the CP channel were replaced (CP mutants) did not show changes in their dissociation constants. The photocycles of these mutants were also examined, and in the case of the EC mutants, the transition to the last step was greatly delayed; on the other hand, in the CP mutants, L2-photointermediate decay was significantly prolonged, except in the case of K215Q, which lacked the O-photointermediate. The importance of Thr218 for binding of Cl(-) to the CP channel was indicated by these results. On the basis of these observations, the possible anion transport mechanism of NpHR was discussed.  相似文献   

11.
Cys126 is a completely conserved residue in triosephosphate isomerase that is proximal to the active site but has been ascribed no specific role in catalysis. A previous study of the C126S and C126A mutants of yeast TIM reported substantial catalytic activity for the mutant enzymes, leading to the suggestion that this residue is implicated in folding and stability [Gonzalez-Mondragon E et al. (2004) Biochemistry 43, 3255-3263]. We re-examined the role of Cys126 with the Plasmodium falciparum enzyme as a model. Five mutants, C126S, C126A, C126V, C126M, and C126T, were characterized. Crystal structures of the 3-phosphoglycolate-bound C126S mutant and the unliganded forms of the C126S and C126A mutants were determined at a resolution of 1.7-2.1 ?. Kinetic studies revealed an approximately five-fold drop in k(cat) for the C126S and C126A mutants, whereas an approximately 10-fold drop was observed for the other three mutants. At ambient temperature, the wild-type enzyme and all five mutants showed no concentration dependence of activity. At higher temperatures (> 40 °C), the mutants showed a significant concentration dependence, with a dramatic loss in activity below 15 μM. The mutants also had diminished thermal stability at low concentration, as monitored by far-UV CD. These results suggest that Cys126 contributes to the stability of the dimer interface through a network of interactions involving His95, Glu97, and Arg98, which form direct contacts across the dimer interface.  相似文献   

12.
Cathepsin E, an intracellular aspartic proteinase of the pepsin family, is composed of two homologous domains, each containing the catalytic Asp residue in a consensus DTG motif. Here we examine the significance of residues in the motifs of rat cathepsin E by substitution of Asp98, Asp283, and Thr284 with other residues using site-directed mutagenesis. Each of the mutant proenzymes, as well as the wild-type protein, was found in culture media and cell extracts when heterologously expressed in human embryonic kidney 293T cells. The single mutants D98A, D283A, and D283E, and the double mutants D98A/D283A and D98E/D283E showed neither autocatalytic processing nor enzymatic activities under acidic conditions. However, the D98E and T284S mutants retained the ability to transform into the mature forms, although they exhibited only about 13 and 40% of the activity of the wild-type enzyme, respectively. The K(m) values of these two mutants were similar to those of the wild-type enzyme, but their k(cat) values were greatly decreased. The K(i) values for pepstatin and the Ascaris pepsin inhibitor of the mutants and the wild-type enzyme were almost the same. The circular dichroism spectra of the two mutants were essentially the same as those of the wild-type enzyme at various pH values. These results indicate that (i) Asp98, Asp283, and Thr284 are indeed critical for catalysis, and (ii) the decrease in the catalytic activity of D98E and T284S mutants is brought about by an effect on the kinetic process from the enzyme-substrate complex to the release of the product.  相似文献   

13.
Random mutagenesis of Thermus thermophilus 3-isopropylmalate dehydrogenase revealed that a substitution of Val126Met in a hinge region caused a marked increase in specific activity, particularly at low temperatures, although the site is far from the binding residues for 3-isopropylmalate and NAD. To understand the molecular mechanism, residue 126 was substituted with one of eight other residues, Gly, Ala, Ser, Thr, Glu, Leu, Ile or Phe. Circular dichroism analyses revealed a decreased thermal stability of the mutants (Delta T ((1/2))= 0-13 degrees C), indicating structural perturbations caused by steric conflict with surrounding residues having larger side chains. Kinetic parameters, k(cat) and K(m) values for isopropylmalate and NAD, were also affected by the mutation, but the resulting k(cat)/K(m) values were similar to that of the wild-type enzyme, suggesting that the change in the catalytic property is caused by the change in free-energy level of the Michaelis complex state relative to that of the initial state. The kinetic parameters and activation enthalpy change (Delta H (double dagger)) showed good correlation with the van der Waals volume of residue 126. These results suggested that the artificial cold adaptation (enhancement of k(cat) value at low temperatures) resulted from the destabilization of the ternary complex caused by the increase in the volume of the residue at position 126.  相似文献   

14.
The salivary agglutinin-interacting adhesin P1 of Streptococcus mutans is anchored to the cell wall via the carboxy (C) terminus, which contains a wall-associated domain, a conserved LPXTGX motif, a hydrophobic domain, and a charged tail. To further investigate the role of the C-terminal anchoring regions in cell wall sorting and anchoring, mutational analysis was performed on P1 in this study. Three truncated P1 mutants and seven site-directed mutants were generated by a polymerase chain reaction-based technique. The mutated P1 genes were returned to the P1-negative S. mutans SM3352 for expression and localization studies by ELISA and Western immunoblotting. The results showed that P1 mutants with deletion of the hydrophobic domain and charged tail, or deletion of the charged tail alone resulted in the secretion of P1 to the culture medium. Results from cellular fractionation experiments with the truncated mutants showed that P1 was not trapped in the membrane or cytoplasm. The site-directed mutants showed normal distribution of P1 to the cell surface as compared to the wild-type. However, when cell walls prepared from the site-directed mutants were boiled with SDS, P1 could be removed readily from the mutants with Thr residue in the LPNTGV motif, altered to either Ser (T1531S) or Phe (T1531F); the mutant with Thr and Gly residues altered to two Phe residues (TG1531-1532FF), and the LPNTGV-deleted mutant (LPNTGV-). In contrast, the wild-type P1 and the other three site-directed P1 mutants (P1529V, N1530I, and G1532F) could not be removed by boiling SDS. When the cell wall P1s from the wild-type, mutants P1529V, N1530I, and G1532F were reacted with an antibody directed against the hydrophobic domain and charged tail, no reaction was detected. However, P1s from mutants T1531S, T1531F, TG1531-1532FF, and LPNTGV- were recognized by the antibody, indicating that the inability of these mutated P1s to firmly link to the cell wall was the result of failure in proteolytic cleavage of the hydrophobic domain and charged tail. In summary, the results suggest that the charged tail plays a decisive role in sorting P1 to the cell surface, while the LPXTGX motif determines the nature of P1-cell wall association. The Thr residue of the LPXTGX motif is required for enzymatic processing to link P1 to the cell wall, presumably via a covalent bond.  相似文献   

15.
Turnover of cyclin E is controlled by SCF(Fbw7). Three isoforms of Fbw7 are produced by alternative splicing. Whereas Fbw7alpha and -gamma are nuclear and the beta-isoform is cytoplasmic in 293T cells, all three isoforms induce cyclin E destruction in an in vivo degradation assay. Cyclin E is phosphorylated on Thr(62), Ser(88), Ser(372), Thr(380), and Ser(384) in vivo. To examine the roles of phosphorylation in cyclin E turnover, a series of alanine point mutations in each of these sites were analyzed for Fbw7-driven degradation. As expected, mutation of the previously characterized residue Thr(380) to alanine led to profound defects of cyclin E turnover, and largely abolished association with Fbw7. Mutation of Thr(62) to alanine led to a dramatic reduction in the extent of Thr(380) phosphorylation, suggesting an indirect effect of this mutation on cyclin E turnover. Nevertheless, phosphopeptides centered at Thr(62) associated with Fbw7, and residual binding of cyclin E(T380A) to Fbw7 was abolished upon mutation of Thr(62), suggesting a minor role for this residue in direct association with Fbw7. Mutation of Ser(384) to alanine also rendered cyclin E resistant to degradation by Fbw7, with the largest effects being observed with Fbw7beta. Cyclin E(S384A) associated more weakly with Fbw7alpha and -beta isoforms but was not defective in Thr(380) phosphorylation. Analysis of the localization of cyclin E mutant proteins indicated selective accumulation of cyclin E(S384A) in the nucleus, which may contribute to the inability of cytoplasmic Fbw7beta to promote turnover of this cyclin E mutant protein.  相似文献   

16.
The mammalian mitochondrial NADP-dependent isocitrate dehydrogenase is a citric acid cycle enzyme and an important contributor to cellular defense against oxidative stress. The Mn(2+)-isocitrate complex of the porcine enzyme was recently crystallized; its structure indicates that Ser(95), Asn(97), and Thr(78) are within hydrogen-bonding distance of the gamma-carboxylate of enzyme-bound isocitrate. We used site-directed mutagenesis to replace each of these residues by Ala and Asp. The wild-type and mutant enzymes were expressed in Escherichia coli and purified to homogeneity. All the enzymes retain their native dimeric structures and secondary structures as monitored by native gel electrophoresis and circular dichroism, respectively. V(max) of the three alanine mutants is decreased to 24%-38% that of wild-type enzyme, with further decreases in the aspartate mutants. For T78A and S95A mutants, the major changes are the 10- to 100-fold increase in the K(m) values for isocitrate and Mn(2+). The results suggest that Thr(78) and Ser(95) function to strengthen the enzyme's affinity for Mn(2+)-isocitrate by hydrogen bonding to the gamma-carboxylate of isocitrate. For the Asn(97) mutants, the K(m) values are much less affected. The major change in the N97A mutant is the increase in pK(a) of the ionizable metal-liganded hydroxyl of enzyme-bound isocitrate from 5.23 in wild type to 6.23 in the mutant enzyme. The hydrogen bond between Asn(97) and the gamma-carboxylate of isocitrate may position the substrate to promote a favorable lowering of the pK of the enzyme-isocitrate complex. Thus, Thr(78), Ser(95), and Asn(97) perform important but distinguishable roles in catalysis by porcine NADP-specific isocitrate dehydrogenase.  相似文献   

17.
To investigate the function of aspartic acid residue 101 and arginine residue 166 in the active site of Escherichia coli alkaline phosphatase (EAP), two single mutants D101S (Asp 101 →Ser) and R166K (Arg 166 →Lys) and a double mutant D101S/R166K of EAP were generated through site-directed mutagenesis based on over-lap PCR method. Their enzymatic kinetic properties, thermal stabilities and possible reaction mechanism were explored. In the presence of inorganic phosphate acceptor, 1 M diethanolamine buffer, the k cat for D101S mutant enzyme increased 10-fold compared to that of wild-type EAP. The mutant R166K has a 2-fold decrease of k cat relative to the wild-type EAP, but the double mutant D101S/R166K was in the middle of them, indicative of an additive effect of these two mutations. On the other hand, the catalytic efficiencies of mutant enzymes are all reduced because of a substantial increase of K m values. All three mutants were more resistant to phosphate inhibitor than the wild-type enzyme. The analysis of the kinetic data suggests that (1) the D101S mutant enzyme obtains a higher catalytic activity by allowing a faster release of the product; (2) the R166K mutant enzyme can reduce the binding of the substrate and phosphate competitive inhibitor; (3) the double mutant enzyme has characteristics of both quicker catalytic turnover number and decreased affinity for competitive inhibitor. Additionally, pre-steady-state kinetics of D101S and D101S/R166K mutants revealed a transient burst followed by a linear steady state phase, obviously different from that of wild-type EAP, suggesting that the rate-limiting step has partially change from the release of phosphate from non-covalent E-Pi complex to the hydrolysis of covalent E-Pi complex for these two mutants.  相似文献   

18.
To investigate the function of aspartic acid residue 101 and arginine residue 166 in the active site of Escherichia coli alkaline phosphatase (EAP), two single mutants D101S (Asp 101 &#77 Ser) and R166K (Arg 166 &#77 Lys) and a double mutant D101S/R166K of EAP were generated through site-directed mutagenesis based on over-lap PCR method. Their enzymatic kinetic properties, thermal stabilities and possible reaction mechanism were explored. In the presence of inorganic phosphate acceptor, 1 M diethanolamine buffer, the k cat for D101S mutant enzyme increased 10-fold compared to that of wild-type EAP. The mutant R166K has a 2-fold decrease of k cat relative to the wild-type EAP, but the double mutant D101S/R166K was in the middle of them, indicative of an additive effect of these two mutations. On the other hand, the catalytic efficiencies of mutant enzymes are all reduced because of a substantial increase of K m values. All three mutants were more resistant to phosphate inhibitor than the wild-type enzyme. The analysis of the kinetic data suggests that (1) the D101S mutant enzyme obtains a higher catalytic activity by allowing a faster release of the product; (2) the R166K mutant enzyme can reduce the binding of the substrate and phosphate competitive inhibitor; (3) the double mutant enzyme has characteristics of both quicker catalytic turnover number and decreased affinity for competitive inhibitor. Additionally, pre-steady-state kinetics of D101S and D101S/R166K mutants revealed a transient burst followed by a linear steady state phase, obviously different from that of wild-type EAP, suggesting that the rate-limiting step has partially change from the release of phosphate from non-covalent E-Pi complex to the hydrolysis of covalent E-Pi complex for these two mutants.  相似文献   

19.
Packing interactions in bacteriophage T4 lysozyme were explored by determining the structural and thermodynamic effects of substitutions for Ala98 and neighboring residues. Ala98 is buried in the core of T4 lysozyme in the interface between two alpha-helices. The Ala98 to Val (A98V) replacement is a temperature-sensitive lesion that lowers the denaturation temperature of the protein by 15 degrees C (pH 3.0, delta delta G = -4.9 kcal/mol) and causes atoms within the two helices to move apart by up to 0.7 A. Additional structural shifts also occur throughout the C-terminal domain. In an attempt to compensate for the A98V replacement, substitutions were made for Val149 and Thr152, which make contact with residue 98. Site-directed mutagenesis was used to construct the multiple mutants A98V/T152S, A98V/V149C/T152S and the control mutants T152S, V149C and A98V/V149I/T152S. These proteins were crystallized, and their high-resolution X-ray crystal structures were determined. None of the second-site substitutions completely alleviates the destabilization or the structural changes caused by A98V. The changes in stability caused by the different mutations are not additive, reflecting both direct interactions between the sites and structural differences among the mutants. As an example, when Thr152 in wild-type lysozyme is replaced with serine, the protein is destabilized by 2.6 kcal/mol. Except for a small movement of Val94 toward the cavity created by removal of the methyl group, the structure of the T152S mutant is very similar to wild-type T4 lysozyme. In contrast, the same Thr152 to Ser replacement in the A98V background causes almost no change in stability. Although the structure of A98V/T152S remains similar to A98V, the combination of T152S with A98V allows relaxation of some of the strain introduced by the Ala98 to Val replacement. These studies show that removal of methyl groups by mutation can be stabilizing (Val98----Ala), neutral (Thr152----Ser in A98V) or destabilizing (Val149----Cys, Thr152----Ser). Such diverse thermodynamic effects are not accounted for by changes in buried surface area or free energies of transfer of wild-type and mutant side-chains. In general, the changes in protein stability caused by a mutation depend not only on changes in the free energy of transfer associated with the substitution, but also on the structural context within which the mutation occurs and on the ability of the surrounding structure to relax in response to the substitution.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

20.
The Saccharomyces cerevisiae pheromone, alpha-factor (WHWLQLKPGQPMY), and Ste2p, its G protein-coupled receptor, were studied as a model for peptide ligand-receptor interaction. The affinities and activities of various synthetic position-10 alpha-factor analogs with Ste2p expressing mutations at residues Ser47 and Thr48 were investigated. All mutant receptors were expressed at a similar level in the cytoplasmic membrane, and their efficacies of signal transduction were similar to that of the wild-type receptor. Mutant receptors differed in binding affinity (Kd) and potency (EC50) for gene induction by alpha-factor. One mutant receptor (S47K,T48K) had dramatically reduced affinity and activity for [Lys10]- and [Orn10]alpha-factor, whereas the affinity for Saccharomyces kluyveri alpha-factor (WHWLSFSKGEPMY) was increased over 20-fold compared with that of wild-type receptor. In contrast, the affinity of [Lys10]- and [Orn10]alpha-factor was increased greatly in a S47E,T48E mutant receptor, whereas the binding of the S. kluyveri alpha-factor was abolished. The affinity of [Lys10]- and [Orn10]alpha-factor for the S47E,T48E receptor dropped 4-6-fold in the presence of 1 m NaCl, whereas the affinity of alpha-factor was not affected by this treatment. These results demonstrate that when bound to its receptor the 10th residue (Gln) of the S. cerevisiae alpha-factor is adjacent to Ser47 and Thr48 residues in the receptor and that the 10th residue of alpha-factors from two Saccharomyces species is responsible for the ligand selectivity to their cognate receptors. Based on these data, we have developed a two-dimensional model of alpha-factor binding to its receptor.  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