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.
Gln34, Gln224, Leu228, and Ser240 are conserved residues in the vicinity of bound IMP in the crystal structure of Escherichia coli adenylosuccinate synthetase. Directed mutations were carried out, and wild-type and mutant enzymes were purified to homogeneity. Circular dichroism spectroscopy indicated no difference in secondary structure between the mutants and the wild-type enzyme in the absence of substrates. Mutants L228A and S240A exhibited modest changes in their initial rate kinetics relative to the wild-type enzyme, suggesting that neither Leu228 nor Ser240 play essential roles in substrate binding or catalysis. The mutants Q224M and Q224E exhibited no significant change in KmGTP and KmASP and modest changes in KmIMP relative to the wild-type enzyme. However, kcat decreased 13-fold for the Q224M mutant and 10(4)-fold for the Q224E mutant relative to the wild-type enzyme. Furthermore, the Q224E mutant showed an optimum pH at 6.2, which is 1.5 pH units lower than that of the wild-type enzyme. Tryptophan emission fluorescence spectra of Q224M, Q224E, and wild-type proteins under denaturing conditions indicate comparable stabilities. Mutant Q34E exhibits a 60-fold decrease in kcat compared with that of the wild-type enzyme, which is attributed to the disruption of the Gln34 to Gln224 hydrogen bond observed in crystal structures. Presented here is a mechanism for the synthetase, whereby Gln224 works in concert with Asp13 to stabilize the 6-oxyanion of IMP.  相似文献   

2.
The molecular processes concomitant with the redox reactions of wild-type and mutant cytochrome c oxidase from Paracoccus denitrificans were analyzed by a combination of protein electrochemistry and Fourier transform infrared (FTIR) difference spectroscopy. Oxidized-minus-reduced FTIR difference spectra in the mid-infrared (4000-1000 cm-1) reflecting full or stepwise oxidation and reduction of the respective cofactor(s) were obtained. In the 1800-1000 cm-1 range, these FTIR difference spectra reflect changes of the polypeptide backbone geometry in the amide I (ca. 1620-1680 cm-1) and amide II (ca. 1560-1540 cm-1) region in response to the redox transition of the cofactor(s). In addition, several modes in the 1600-1200 cm-1 range can be tentatively attributed to heme modes. A peak at 1746 cm-1 associated with the oxidized form and a peak at 1734 cm-1 associated with the reduced form were previously discussed by us as proton transfer between Asp or Glu side chain modes in the course of the redox reaction of the enzyme [Hellwig, P., Rost, B., Kaiser, U., Ostermeier, C., Michel, H., and M?ntele, W. (1996) FEBS Lett. 385, 53-57]. These signals were resolved into several components associated with the oxidation of different cofactors. For a stepwise potential titration from the fully reduced state (-0.5 V) to the fully oxidized state (+0.5 V), a small component at 1738 cm-1 develops in the potential range of approximately +0.15 V and disappears at more positive potentials while the main component at 1746 cm-1 appears in the range of approximately +0.20 V (all potentials quoted vs Ag/AgCl/3 M KCl). This observation clearly indicates two different ionizable residues involved in redox-induced proton transfer. The major component at 1746 cm-1 is completely lost in the FTIR difference spectra of the Glu 278 Gln mutant enzyme. In the spectrum of the subunit I Glu 278 Asp mutant enzyme, the major component of the discussed difference band is lost. In contrast, the complete difference signal of the wild-type enzyme is preserved in the Asp 124 Asn, Asp 124 Ser, and Asp 399 Asn variants, which are critical residues in the discussed proton pump channel as suggested from structure and mutagenesis experiments. On the basis of these difference spectra of mutants, we present further evidence that glutamic acid 278 in subunit I is a crucial residue for the redox reaction. Potential titrations performed simultaneously for the IR and for the UV/VIS indicate that the signal related to Glu 278 is coupled to the electron transfer to/from heme a; however, additional involvement of CuB electron transfer cannot be excluded.  相似文献   

3.
AMPA receptor GluRA subunits with mutations at position 750, a residue shown previously to control allosteric regulation by cyclothiazide, were analyzed for modulation of deactivation and desensitization by cyclothiazide, aniracetam, and thiocyanate. Point mutations from Ser to Asn, Ala, Asp, Gly, Gln, Met, Cys, Thr, Leu, Val, and Tyr were constructed in GluRAflip. The last four of these mutants were not functional; S750D was active only in the presence of cyclothiazide, and the remaining mutants exhibited altered rates of deactivation and desensitization for control responses to glutamate, and showed differential modulation by cyclothiazide and aniracetam. Results from kinetic analysis are consistent with aniracetam and cyclothiazide acting via distinct mechanisms. Our experiments demonstrate for the first time the functional importance of residue 750 in regulating intrinsic channel-gating kinetics and emphasize the biological significance of alternative splicing in the M3-M4 extracellular loop.  相似文献   

4.
RGS proteins (regulators of G protein signaling) are GTPase activating proteins (GAPs) for Gi and Gq families of heterotrimeric G proteins but have not been found to interact with Gs alpha. The Gs alpha residue Asp229 has been suggested to be responsible for the inability of RGS proteins to interact with Gs alpha [Natochin, M., and Artemyev, N. O. (1998) J. Biol. Chem. 273, 4300-4303]. To test this hypothesis, we have investigated the possibility of generating an interaction between Gs alpha and RGS proteins by substituting Gs alpha Asp229 with Ser and replacing the potential Gs alpha Asp229 contact residues in RGS16, Glu129 and Asn131, by Ala and Ser, respectively. RGS16 and its mutants failed to interact with Gs alpha. A single mutation of Gs alpha, Asp229Ser, rendered the Gs alpha subunit with the ability to interact with RGS16 and RGS4. Like RGS protein binding to Gi and Gq alpha-subunits, RGS16 preferentially recognized the AlF4--bound conformation of Gs alpha Asp229Ser. In a single-turnover assay, RGS16 maximally stimulated GTPase activity of Gs alpha Asp229Ser by approximately 5-fold with an EC50 value of 7.5 microM. Our findings demonstrate that Asp229 of Gs alpha represents a major barrier for Gs alpha interaction with known RGS proteins.  相似文献   

5.
Mycoplasma hominis mutants were selected stepwise for resistance to ofloxacin and sparfloxacin, and their gyrA, gyrB, parC, and parE quinolone resistance-determining regions were characterized. For ofloxacin, four rounds of selection yielded six first-, six second-, five third-, and two fourth-step mutants. The first-step mutants harbored a single Asp426-->Asn substitution in ParE. GyrA changes (Ser83-->Leu or Trp) were found only from the third round of selection. With sparfloxacin, three rounds of selection generated 4 first-, 7 second-, and 10 third-step mutants. In contrast to ofloxacin resistance, GyrA mutations (Ser83-->Leu or Ser84-->Trp) were detected in the first-step mutants prior to ParC changes (Glu84-->Lys), which appeared only after the second round of selection. Further analysis of eight multistep-selected mutants of M. hominis that were previously described (2) revealed that they carried mutations in ParE (Asp426-->Asn), GyrA (Ser83-->Leu) and ParE (Asp426-->Asn), GyrA (Ser83-->Leu) and ParC (Ser80-->Ile), or ParC (Ser80-->Ile) alone, depending on the fluoroquinolone used for selection, i.e., ciprofloxacin, norfloxacin, ofloxacin, or pefloxacin, respectively. These data indicate that in M. hominis DNA gyrase is the primary target of sparfloxacin whereas topoisomerase IV is the primary target of pefloxacin, ofloxacin, and ciprofloxacin.  相似文献   

6.
Glutamine 105 in the third transmembrane helix of the thyrotropin-releasing hormone receptor (TRH-R) occupies a position equivalent to a conserved negatively charged residue in receptors for biogenic amines where it acts as counterion interacting with the cationic amine moiety of the ligand. Maximum levels of response to TRH in oocytes expressing wild-type TRH-Rs were indistinguishable from those of oocytes expressing receptors mutated to Glu, Asn, or Asp in position 105. However, the EC50 values for activation of oocyte responses increased more than 500 times in oocytes expressing mutant Glu105 receptors, in which the amido group of Gln105 has been removed by site-directed mutagenesis. Charge effects do not seem to be involved in the huge effect of mutating Gln105 to Glu, since mutation of Gln105 to Asp induces only a 15-fold increase in EC50. Furthermore, no change in EC50 is observed after mutation of Asn110 to Asp. The affinity shift (identified by changes in EC50 values for systems of comparable efficacy) in Glu105 mutant receptors was partially recovered in oocytes expressing Asn105 mutant receptors. These results and those obtained after substitution of Lys, Leu, Tyr, and Ser for Gln105 suggest that the presence and the correct position of the Gln hydrogen bond-donor amido group are important for normal functionality of the receptor. In wild type or Asp105 mutant receptors showing the same maximal responses, decreases in affinity with TRH and methyl-histidyl-TRH correlated with increased dissociation rates of hormone from the receptor. Rapid dilution experiments following subsecond stimulation indicate that the TRH-R is converted rapidly from a form showing fast dissociation kinetics to a form from which the hormone dissociates slowly. Mutation of residue 105 impairs the receptor shift between these two forms. This effect was demonstrated in a direct way by comparing [3H]methyl-histidyl-TRH dissociation rates in COS-7 cells transfected with either wild type or Asp105 mutant TRH-Rs. Thus, residues located in transmembrane helix III positions equivalent to those of the counterions for biogenic amines, regulate hormone-receptor interactions in the TRH receptor (and perhaps other receptors). Furthermore, the nature of the amino acid in these positions may also play a role, directly or indirectly, in conformational changes leading to receptor activation, and hence to signal transduction.  相似文献   

7.
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   

8.
We study, by flash kinetic spectrophotometry on the microsecond time scale, the effects of ionic strength and viscosity on the kinetics of oxidative quenching of the triplet state of zinc cytochrome c (3Zncyt) by the wild-type form and the following nine mutants of cupriplastocyanin: Leu12Glu, Leu12Asn, Phe35Tyr, Gln88Glu, Tyr83Phe, Tyr83His, Asp42Asn, Glu43Asn, and the double mutant Glu59Lys/Glu60Gln. The unimolecular rate constants for the quenching reactions within the persistent diprotein complex, which predominates at low ionic strengths, and within the transient diprotein complex, which is involved at higher ionic strengths, are equal irrespective of the mutation. Evidently, the two complexes are the same. In both reactions, the rate-limiting step is rearrangement of the diprotein complex from a configuration optimal for docking to the one optimal for the subsequent electron-transfer step, which is fast. We investigate the effects of plastocyanin mutations on this rearrangement, which gates the overall electron-transfer reaction. Conversion of the carboxylate anions into amide groups in the lower acidic cluster (residues 42 and 43), replacement of Tyr83 with other aromatic residues, and mutations in the hydrophobic patch in plastocyanin do not significantly affect the rearrangement. Conversion of a pair of carboxylate anions into a cationic and a neutral residue in the upper acidic cluster (residues 59 and 60) impedes the rearrangement. Creation of an anion at position 88, between the upper acidic cluster and the hydrophobic patch, facilitates the rearrangement. The rate constant for the rearrangement smoothly decreases as the solution viscosity increases, irrespective of the mutation. Fittings of this dependence to the modified Kramers's equation and to an empirical equation show that zinc cytochrome c follows the same trajectory on the surfaces of all the plastocyanin mutants but that the obstacles along the way vary as mutations alter the electrostatic potential. Mutations that affect protein association (i.e., change the binding constant) do not necessarily affect the reaction between the associated proteins (i.e., the rate constant) and vice versa. All of the kinetic and thermodynamic effects and noneffects of mutations consistently indicate that in the protein rearrangement the basic patch of zinc cytochrome c moves from a position between the two acidic clusters to a position at or near the upper acidic cluster.  相似文献   

9.
Heme peroxidases of prokaryotic, plant and fungal origin share the essential His and Arg catalytic residues of the distal cavity and a proximal His bound to heme iron. Spectroscopic techniques, in contrast to X-ray crystallography, are well suited to detect the precise structure, spin and coordination states of the heme as influenced by its near environment. Resonance Raman and electronic absorption spectra obtained at various pH values for Fe3+ and Fe2+ forms of distal Arg51 mutants of the fungal Coprinus cinereus peroxidase are reported, together with the fluoride adducts at pH 5.0. This basic catalytic residue has been replaced by the aliphatic residue Leu, the polar residues Asn and Gln and the basic residue Lys (Arg51-->Leu, Asn, Gln, and Lys, respectively). These mutations cause changes in the coordination and spin states of the heme iron, and in the v(Fe-Im) stretching frequency. The variations are explained in terms of pH-dependent changes, charge location, size and hydrogen-bonding acceptor/donor properties of the residue at position 51. The present work indicates that the hydrogen-bond capability of the residue in position 51 influences the occupancy of water molecules in the distal cavity and the ability to form stable complexes between anionic ligands and the heme Fe atom.  相似文献   

10.
Histidine-containing protein (HPr) is a phosphocarrier protein of the bacterial phosphoenolpyruvate:sugar phosphotransferase system. HPr is phosphorylated at the active site residue, His15, by phosphoenolpyruvate-dependent enzyme I in the first enzyme reaction in the process of phosphoryl transfer to sugar. In many Gram-positive bacterial species HPr may also be phosphorylated at Ser46 by an ATP-dependent protein kinase but not in the Gram-negative Escherichia coli and Salmonella typhimurium. One effect of the phosphorylation at Ser46 is to make HPr a poor acceptor for phosphorylation at His15. In Bacillus subtilis HPr, the mutation Ser46Asp mimics the effects of phosphorylation. A series of mutations were made at Ser46 in E. coli HPr: Ala, Arg, Asn, Asp, Glu, and Gly. The two acidic replacements mimic the effects of phosphorylation of Ser46 in HPrs from Gram-positive bacteria. In particular, when mutated to Asp46, the His 15 phosphoacceptor activity (enzyme I Km/Kcat) decreases by about 2000-fold (enzyme I Km, 4 mM HPr; Kcat, approximately 30%). The alanine and glycine mutations had near-wild-type properties, and the asparagine and arginine mutations yielded small changes to the Km values. The crystallographic tertiary structure of Ser46Asp HPr has been determined at 1.5 A resolution, and several changes have been observed which appear to be the effect of the mutation. There is a tightening of helix B, which is demonstrated by a consistent shortening of hydrogen bond lengths throughout the helix as compared to the wild-type structure. There is a repositioning of the Gly54 residue to adopt a 3(10) helical pattern which is not present in the wild-type HPr. In addition, the higher resolution of the mutant structure allows for a more definitive placement of the carbonyl of Pro11. The consequence of this change is that there is no torsion angle strain at residue 16. This result suggests that there is no active site torsion angle strain in wild-type E. coli HPr. The lack of substantial change at the active center of E. coli HPr Ser46Asp HPr suggests that the effect of the Ser46 phosphorylation in HPrs from Gram-positive bacteria is due to an electrostatic interference with enzyme I binding.  相似文献   

11.
In the N-terminal region of the alpha-helix of the c-type lysozymes, two Asx residues exist at the 18th and 27th positions. Hen lysozyme has Asp18/Asn27 (18D/27N), and we prepared three mutant lysozymes, Asn18/Asn27 (18N/27N), Asn18/Asp27 (18N/27D), and Asp18/Asp27 (18D/27D). The stability of the wild-type (18D/27N) lysozyme supported the existence of a hydrogen bond between the side chain of Asp18 and the amide group at the N1 position in the alpha-helix, while the stability of the 18N/27D lysozyme supported the presence of the capping box between the Ser24 (N-cap) and Asp27 residues. Although electrostatic repulsion was observed between Asp18 and Asp27 residues in 18D/27D lysozyme, the dissociation of each residue contributed to stabilizing the B-helix in 18D/27D lysozyme through hydrogen bonding and charge-helix macrodipole interaction. This is the first evidence that two neighboring negative charges at the N-terminus of the helix both increased the stability of the protein.  相似文献   

12.
The stoichiometry of the interaction between Erythrina variegata chymotrypsin inhibitor ECI and chymotrypsin was reinvestigated by analysis of their complex with ultracentrifugation and with amino acid analysis of the components separated. The amino acid analysis clearly showed that the stoichiometry of ECI and chymotrypsin was 1:1, though the apparent molecular mass of the complex was estimated to be 60 kDa. To examine the contribution of Leu64 (the P1 residue) to the inhibitory activity of ECI, a complete set of mutated inhibitors in which the amino acid at position 64 was replaced by 19 other amino acid residues was constructed by means of site-directed mutagenesis. Potent inhibitory activities (Ki, 1.3-4.6 x 10(-8) M) exceeding that of the wild-type ECI (Ki, 9.8 x 10(-8) M) were present in the mutant proteins L64F, L64M, L64W, and L64Y. The inhibitory activity of the mutant L64R was practically identical to that of the wild-type ECI. All other mutants exhibited slightly decreased inhibitory activities with Ki values of 1.9-4.6 x 10(-7) M. These results indicate that ECI-chymotrypsin interaction involves not only the P1 site residue but also other residue(s) of ECI. A series of individual alanine mutations was then constructed in residues Gln62 (P3), Phe63 (P2), Ser65 (P1'), Thr66 (P2'), and Phe67 (P3') in order to evaluate the contribution of each residue in the primary binding loop to the inhibitory activity. Replacement of Gln62, Phe63, and Phe67 with Ala residues decreased the inhibitory activity, the Ki values being increased by approximately 3-4-fold; but replacement of Ser65 and Thr66 had relatively little effect. This suggests that the P2, P3, and P3' residues, together with the P1 residue, in the primary binding loop play an important role in the inhibitory activity toward chymotrypsin.  相似文献   

13.
p53 is very often mutated in human cancers. The majority of alterations are missense mutations located within the DNA-binding domain of the protein. Many reports have characterized such mutant proteins. Little is known, however, about the properties of proteins that have a missense mutation outside this domain. We investigated here the properties of 8 mutant proteins identified in human tumors as having a missense mutation in the tetramerization domain. The Arg342Gln, Glu349Asp and Gln354Arg proteins behaved like wild-type both in vitro and in cells. Two mutants, Arg342Pro and Leu344Pro, were inactive in all assays. Finally, the 3 mutant proteins Leu330His, Arg337Cys and Arg337Leu, which are inactive in vitro, showed no activity at low expression levels in cells but became active at higher expression levels. Our results reveal new phenotypes for p53 mutants and suggest that sequencing of the p53 gene from patients with tumors should be extended to exons 9 and 10 in clinical investigations.  相似文献   

14.
Two-component signaling systems are used by bacteria, plants, and lower eukaryotes to adapt to environmental changes. The first component, a protein kinase, responds to a signal by phosphorylating the second component; a response regulator protein that often acts by inducing the expression of specific genes. Response regulators also have an autophosphatase activity that ensures that the proteins are not permanently activated by phosphorylation. The magnitude of this activity varies by at least 1000-fold between various response regulators, and the molecular features responsible for this varied autophosphatase activity have not been clearly defined. Using wild-type and mutant derivatives of the sporulation response regulator Spo0F, it has been demonstrated that a key residue in determining the magnitude of this activity is that at position 56 of Spo0F approximately P; this residue is adjacent to the site of phosphorylation, Asp 54. For example, Spo0F approximately P K56N has a 23-fold greater autophosphatase activity (t1/2 = 8 min) than wild-type Spo0F approximately P (t1/2 = 180 min). It is suggested that, by analogy to the GTPase activity of p21(ras) and by examining the crystallographic structure of Spo0F, that the carboxyamide of the mutant Asn 56 may favorably position a catalytic water near the protein acyl phosphate to promote Spo0F approximately P K56N hydrolysis. It is also deduced that Lys 56 in the wild-type protein is critical for the efficient interaction and phosphoryl transfer between Spo0F and it's cognate protein kinase, KinA. Comparison of the known response regulators shows that inefficient autophosphatases (t1/2 on the order of hours) typically contain an amino acid residue with a long side chain at the position equivalent to 56 in Spo0F, whereas efficient autophosphatases (t1/2 on the order of minutes) frequently contain a residue with a carboxyamide or carboxylate side chain at this position. It appears that, by altering residues adjacent to the active site, the autophosphatase activity of response regulator proteins has been attenuated to match the diverse biological roles played by these proteins.  相似文献   

15.
The average globular protein contains 30% alpha-helix, the most common type of secondary structure. Some amino acids occur more frequently in alpha-helices than others; this tendency is known as helix propensity. Here we derive a helix propensity scale for solvent-exposed residues in the middle positions of alpha-helices. The scale is based on measurements of helix propensity in 11 systems, including both proteins and peptides. Alanine has the highest helix propensity, and, excluding proline, glycine has the lowest, approximately 1 kcal/mol less favorable than alanine. Based on our analysis, the helix propensities of the amino acids are as follows (kcal/mol): Ala = 0, Leu = 0.21, Arg = 0.21, Met = 0.24, Lys = 0.26, Gln = 0.39, Glu = 0.40, Ile = 0.41, Trp = 0.49, Ser = 0.50, Tyr = 0. 53, Phe = 0.54, Val = 0.61, His = 0.61, Asn = 0.65, Thr = 0.66, Cys = 0.68, Asp = 0.69, and Gly = 1.  相似文献   

16.
The crystal structure of a lipase from the bacterium Chromobacterium viscosum ATCC 6918 (CVL) has been determined by isomorphous replacement and refined at 1.6 angstroms resolution to an R-factor of 17.8%. The lipase has the overall topology of an alpha/beta type protein, which was also found for previously determined lipase structures. The catalytic triad of the active center consists of the residues Ser87, Asp263 and His285. These residues are not exposed to the solvent, but a narrow channel connects them with the molecular surface. This conformation is very similar to the previously reported closed conformation of Pseudomonas glumae lipase (PGL), but superposition of the two lipase structures reveals several conformational differences. r.m.s. deviations greater than 2 angstroms are found for the C alpha-atoms of the polypeptide chains from His15 to Asp28, from Leu49 to Ser54 and from Lys128 to Gln158. Compared to the PGL structure in the CVL structure, three alpha-helical fragments are shorter, one beta-strand is longer and an additional antiparallel beta-sheet is found. In contrast to PGL, CVL displays an oxyanion hole, which is stabilized by the amide nitrogen atoms of Leu17 and Gln88, and a cis-peptide bond between Gln291 and Leu292. CVL contains a Ca2+, like the PGL, which is coordinated by four oxygen atoms from the protein and two water molecules.  相似文献   

17.
The photovoltage kinetics of the bacteriorhodopsin mutants Asp212-->Asn and Asp85-->Asn after excitation at 580 nm have been investigated in the pH range from 0 to 11. With the mutant Asp85-->Asn (D85N) at pH 7 no net charge translocation is observed and the signal is the same, both in the presence of Cl- (150 mM) and in its absence (75 mM SO4(2-)). Under both conditions the color of the pigment is blue (lambda max = 615 nm). The time course of the photovoltage kinetics is similar to that of the acid-blue form of wild-type, except that an additional transient charge motion occurs with time constants of 60 microseconds and 1.3 ms, indicating the transient deprotonation and reprotonation of an unknown group to and from the extracellular side of the membrane. It is suggested that this is the group XH, which is responsible for proton release in wild-type. At pH 1, the photovoltage signal of D85N changes upon the addition of Cl- from that characteristic for the acid-blue state of wild-type to that characteristic for the acid-purple state. Therefore, the protonation of the group at position at 85 is necessary, but not sufficient for the chloride-binding. At pH 11, well above the pKa of the Schiff base, there is a mixture of "M-like" and "N-like" states. Net proton transport in the same direction as in wild-type is restored in D85N from this N-like state. With the mutant Asp212-->Asn (D212N), time-resolved photovoltage measurements show that in the absence of halide ions the signal is similar to that of the acid-blue form of wild-type and that no net charge translocation occurs in the entire pH range from 0 to 11. Upon addition of Cl- in the pH range from 3.8 to 7.2 the color of the pigment returns to purple and the photovoltage experiments indicate that net proton pumping is restored. However, this Cl(-)-induced activation of net charge-transport in D212N is only partial. Outside this pH range, no net charge transport is observed even in the presence of chloride, and the photovoltage shows the same chloride-dependent features as those accompanying the acid-blue to acid-purple transition of the wild-type.  相似文献   

18.
We have developed a useful strategy for identifying amino acid spin systems and side-chain carbon resonance assignments in small 15N-, 13C-enriched proteins. Multidimensional constant-time pulsed field gradient (PFG) HCC(CO)NH-TOCSY experiments provide side-chain resonance frequency information and establish connectivities between sequential amino acid spin systems. In PFG HCC(CO)NH-TOCSY experiments recorded with a properly tuned constant-time period for frequency labeling of aliphatic 13C resonances, phases of cross peaks provide information that is useful for identifying spin system types. When combined with 13C chemical shift information, these patterns allow identification of the following spin system types: Gly, Ala, Thr, Val, Leu, Ile, Lys, Arg, Pro, long-type (i.e., Gln, Glu and Met), Ser, and AMX-type (i.e., Asp, Asn, Cys, His, Phe, Trp and Tyr).  相似文献   

19.
Mechanism-based inactivators of beta-lactamases are used to overcome the resistance of clinical pathogens to beta-lactam antibiotics. This strategy can itself be overcome by mutations of the beta-lactamase that compromise the effectiveness of their inactivation. We used PCR mutagenesis of the TEM-1 beta-lactamase gene and sequenced the genes of 20 mutants that grew in the presence of ampicillin-clavulanate. Eleven different mutant genes from these strains contained from 1 to 10 mutations. Each had a replacement of one of the four residues, Met69, Ser130, Arg244, and Asn276, whose substitutions by themselves had been shown to result in inhibitor resistance. None of the mutant enzymes with multiple amino acid substitutions generated in this study conferred higher levels of resistance to ampicillin alone or ampicillin with beta-lactamase inactivators (clavulanate, sulbactam, or tazobactam) than the levels of resistance conferred by the corresponding single-mutant enzymes. Of the four enzymes with just a single mutation (Ser130Gly, Arg244Cys, Arg244Ser, or Asn276Asp), the Asn276Asp beta-lactamase conferred a wild-type level of ampicillin resistance and the highest levels of resistance to ampicillin in the presence of inhibitors. Site-directed random mutagenesis of the Ser130 codon yielded no other mutant with replacement of Ser130 besides Ser130Gly that produced ampicillin-clavulanate resistance. Thus, despite PCR mutagenesis we found no new mutant TEM beta-lactamase that conferred a level of resistance to ampicillin plus inactivators greater than that produced by the single-mutation enzymes that have already been reported in clinical isolates. Although this is reassuring, one must caution that other combinations of multiple mutations might still produce unexpected resistance.  相似文献   

20.
Glu-198 of human matrilysin is a conserved residue in the matrix metalloproteinases and is considered to play an important role in catalysis by acting as a general base catalyst toward the zinc-bound water molecule, on the basis of mechanistic proposals for other zinc proteases. In the present study, Glu-198 is mutated into Asp, Cys, Gln, and Ala, and the zinc binding properties, kinetic parameters, and pH dependence of each mutant are determined in order to examine the role of Glu-198 in catalysis. The mutations chosen either modify (C and D) or eliminate (A and Q) the general base properties of residue-198. All the mutants bind 2 mol of zinc per mol of enzyme, indicating that Glu-198 is not crucial to the binding of the catalytic zinc to the enzyme. The value of kcat/Km for the E198D mutant is only 4-fold lower than that of wild-type enzyme at the pH optimum of 7.5, while that for the E198C mutant is decreased by 160-fold. The E198Q and E198A enzymes containing the mutations that have eliminated the nucleophilic and acid/base properties of the residue are still active, having lower kcat/Km values of 590- and 1900-fold, respectively. The decrease in activity of all the mutants is essentially due to a decrease in kcat. The kcat/Km values of the mutants as a function of pH display broad bell-shaped curves that are similar to the wild-type enzyme. The acidic pKa value is not greatly affected by the change in the chemical properties of residue-198. The similarity in the pH profiles for the mutant and wild-type enzymes indicates that the ionization of Glu-198 is not responsible for the acidic pKa. Ionization of the zinc-bound water may be responsible for this pKa since the three His ligands and the scaffolding of the matrilysin catalytic zinc site are different from that observed in carboxypeptidase A and would predict a lower pKa for the metal-bound water. If the zinc-bound water is the nucleophile in the reaction, the role of Glu-198 in catalysis may be to stabilize the transition state or act as a general acid catalyst after the rate-determining step.  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